content
stringlengths 1
15.9M
|
---|
\section{Monopole condensation and chiral symmetry breaking}
Recently,
chiral symmetry breaking
in the maximal abelian field and in its monopole-part
has been examined
on SU(2) lattice at finite temperature by one of the present authors.
\cite{Miya}
It has been found that
chiral condensates (in quench level) in the U$_1$ field
and in its monopole part ,
keep the same feature of chiral breaking
with that observed in original SU(2) gauge field.
Existence of light pion in the maximal abelian field has also
been reported at zero temperature.\cite{Wolo}
As for confinement, condensation of monopoles has got
increasing evidences.
Wriggling of magnetic current in confinement phase\cite{Kron}
abelian dominance in confinement,\cite{Brands,Hioki}
saturation of area law of Wilson loops by monopole
contribution,\cite{Ohno}
evidence of monopole condensation by
entropy dominance over energy of magnetic current \cite{Shiba}
and finiteness of an order parameter in confinement phase\cite{Giacomo}
have been reported.
Those observations give a support for an idea that
monopole condensation is responsible for chiral symmetry breaking
as well as confinement.
Historically, such idea has been
discussed several times. Banks and Casher first arged that
confinement breaks chiral symmetry in 1980.\cite{Casher}
Later, some authors tried to show chiral symmetry breaking
based on dual Ginzburg-Landau theory.\cite{Baker,Klein,Kamizawa}
A recent work by Suganuma, Sasaki and Toki
has given an affirmative answer for this possibility.\cite{Suga}
Here we investigate pseudoscalar boson in the field dominated
by monopoles on lattice for a further investigation.
We study relationship between instantons and monopoles
also by observing topological charge.
Present analyses have been done
using Intel Paragon
XP/S(56nodes) at the Institute for
Numerical Simulations and Applied Mathematics in Hiroshima University.
\section{ U$_1$ gauge field by maximal abelian projection and
its decomposition into monopole and photon parts}
{}From an SU(2) gauge field, we extract U$_1$ field through maximal abelian
projection\cite{tHooft2}
by maximizing
$$
R= \sum _{n,\mu} Tr(\sigma_3U_{\mu}(n)
\sigma_3U_{\mu}(n)^{\dagger}) , \eqno[1]
$$
and spliting SU(2) gauge link $U_{\mu}(n)$ as
$$
U_{\mu}(n) = \left[\matrix{r_c(n) & -c_{\mu}(n)
\cr c_{\mu}(n)^* &
r_c(n) }\right ] \times
\left[ \matrix{ e^{i\theta^{}_{\mu}(n)} & 0 \cr 0 & e^{-i\theta^{}_{\mu}(n)}
} \right ] , \eqno[2]
$$
where $r_c(n)=\sqrt{1-|c_{\mu}(n)|^2}$.
Here, $\theta^{}_{\mu}(n)$ is the U$_1$ gauge field and
$c_{\mu}(n)$ is interpreted as a charged matter field under the U$_1$
gauge transformation.
Decomposition into the singular(monopole) and the regular(photon)
parts is carried out following Matsubara et al.
\cite{Matsu1,Suzuki}
On lattice , every quantity is regular at finite lattice spacing and we
can work in Landau gauge. Then we have Landau gauge field
as
$$
\theta^L_{\mu}(n)=\sum _m G(n-m) \partial_{\lambda}
\theta^{}_{\lambda\mu}(m) , \eqno[3]
$$
where the abelian field strength $\theta^{}_{\mu\nu}$ is given by
$
\theta^{}_{\mu\nu} =
\partial_{\mu}\theta^{}_{\nu}-\partial_{\nu}\theta^{}_{\mu}$
and $G(n)$ is lattice Coulomb propagator and $\partial$ is
derivative on lattice.
Here we decompose field strength following
DeGrand-Toussaint as,\cite{DeG}
$$
\theta^{}_{\mu\nu}=\bar \theta^{}_{\mu\nu} + 2\pi M_{\mu\nu}
, \eqno[4]
$$
where $ -\pi < \bar \theta^{}_{\mu\nu} < \pi$ and
$M_{\mu\nu}$ is Dirac string.
Substituting eq.[4] into eq.[3], we have a decomposition of
$\theta^L_{\mu}(n)$
into a regular part $\theta_{\mu}^{Ph}(n)$ and a singular
part $\theta_{\mu}^{Ds}(n)$
where
$$
\theta_{\mu}^{Ph}(n) = \sum _m G(n-m) \partial_{\lambda}
\bar \theta^{}_{\lambda\mu}(m) , \eqno[5]
$$
and
$$
\theta_{\mu}^{Ds}(n) = 2\pi \sum _m G(n-m) \partial_{\lambda}
M_{\lambda\mu}(m) . \eqno[6]
$$
It is noted that $\theta_{\mu}^L(n)=\theta_{\mu}^{Ph}(n)+\theta_{\mu}^{Ds}(n)$.
Such decomposition has been discussed
by de Forcrand et al.\cite{Forc}
As a result of the decomposition, the singular part carrys almost
equal number of magnetic currents ($95 - 98\%$) to
that in the $U_1$ field whereas
the number of electric currents is less than
a few $\%$ of that in the $U_1$
field. On the other hand, situation is
just opposite in the regular part.\cite{Miya}
By this reason,
the singular part is called as monopole part and
the regular part as photon part.
\section{ Pseudoscalar correlator in the field dominated by monopoles}
Existence of light pseudoscalar boson is examined to confirm chiral broken
phase in U$_1$ and the singular gauge field $\theta_\mu (n)^{Ds}$.
We measure pseudo
scalar correlator by staggered quark on $16^3\times 32$ lattice
at $\beta=2.2$.
The quark operator is given by
$$
D=-ma\delta_{n,m}-\sum_{\mu}\eta (n)[ V_{\mu}(n)\delta_{n+\mu,m}-
V_{\mu}^{\dagger}(n-\mu)\delta_{n+\mu,m}], \eqno[7]
$$
and
$V_{\mu}(n)$ is set to be either $U_{\mu}(n)$ or
$e^{i\theta_{\mu}(n)}$ or
$e^{i\theta_{\mu}^{Ds}(n)}$ or
$e^{i\theta_{\mu}^{Ph}(n)}$.
Quark mass parameter is taken as $ma=0.005$
and number of configurations is 35.
Results are presented in Fig.1.
The pion correlators in SU(2), U$_1$ and
$\theta_\mu (n)^{Ds}$ gauge fields have similar shape.
Local masses in those correlators take similar values .
They are
$$
am(0^-)=0.174(38) , \ \ 0.169(61) , \ \ 0.162(82) , \eqno[8].
$$
in SU(2), U$_1$ and the monopole part, respectively, at $t=6 \sim 10$.
Although chiral limit ($ m \rightarrow 0 $) has not been examined,
existence of light pseudoscalar meson in those gauge field is confirmed.
On the other hand,
that in the regular part $\theta_\mu (n)^{Ph}$ behave differently as shown in
Fig.2 . In order to understand this behavior,
we present pseudoscalar correlator of free-staggered fermion in Fig.3 .
Saw tooth like behavior of the correlator in the regular gauge field is similar
to that of free staggered quark. Therefore
in the regular gauge field, there seems no light pion and no evidence
for chiral symmetry breaking.
\section{Topological charge in the photon and the monopole part}
Here we report an evidence of strong correlation between instantons
and monopoles. Relationship between instantons in nonabelian gauge
field and monopoles in its $U_1$ sector has not well clarified.
Nevertheless , intimate relation is expected due to several reasons.
Both are non-trivial topological objects. Instanton is known to couple
with chiral symmetry breaking. Actually, we have shown
that monopole seems
relevant to the breaking in the previous work \cite{Miya}
and in this paper.
Quantities examined here are charge $Q$, abolute integral of topological
charge density $I_Q$ and a normalized action $\tilde s$
defined by
$$
Q={1 \over 32\pi^2} \sum \epsilon_{\mu \nu \lambda \sigma}
tr[ P_{\mu \nu}P_{\lambda \sigma}] , \eqno[9]
$$
$$
I_Q={1 \over 32\pi^2} \sum |\epsilon_{\mu \nu \lambda \sigma}
tr[ P_{\mu \nu}P_{\lambda \sigma}]| , \eqno[10]
$$
$$
\tilde s= {1 \over 8\pi^2 } \sum (1-{1 \over 2} tr[P_{\mu \nu}]) ,
\eqno[11]
$$
where $P_{\mu \nu}$ is a plaquette in $\mu \nu$ -plane.
We evaluate those quantities for the Ph-part and the Ds-part as well as
original SU(2) link variables with using cooling method.
Since, $Q$ and $I_Q$ are those of SU(2) object, we reconstruct
SU(2) link variables $U_{\mu}^{Ph}(n)$ and $U_{\mu}^{Ds}(n)$
from $\theta^{Ph}_{\mu}(n)$ and $\theta^{Ds}_{\mu}(n)$ by multipling
the matter part $M_{\mu}(n)$
in the following way;
$$
U_{\mu}^{i}(n) = \left[\matrix{r_c(n) & -c_{\mu}(n)
\cr c_{\mu}(n)^* &
r_c(n) }\right ] \times
\left[ \matrix{ e^{i\theta^{i}_{\mu}(n)} & 0 \cr 0 &
e^{-i\theta^{i}_{\mu}(n)} } \right ] , \eqno[11]
$$
where $i$ takes either Ph or Ds .
An sample of 40 configurations, we cool those three fields, $U_{\mu}(n)$ ,
$U_{\mu}^{Ph}(n)$,$U_{\mu}^{Ds}(n)$ and monitor $Q$ , $I_Q$ and $\tilde s$.
Fig.4a,4b and 4c show typical examples of cooling curves.
Although there is
quantitative difference, $Q$ , $I_Q$ and $\tilde s$ of
SU(2) and Ds-part have nontrivial plateaus while those of Ph-part die
very quickly.
The trivial feature in Ph-part is common in all 40 configurations.
Fig.5a and 5b show correlation between topological charges of SU(2)
and
Ph-part or Ds-part at 3 cooling.
Fig.s 6a and 6b are similar one but that at 100 cooling.
We see that a strong correlation between SU(2) and Ds-part while Ph-part
is trivial and no correlation. Similar features are observed also in $I_Q$
and $\tilde s$. These suggest that Ds-part, i.e. monopole part ,
is responsible to topological charge and has close connection to instantons.
\section{Summary and conclusion}
Pseudoscalar correlator is examined in the maximal projected U$_1$ gauge field
and in its decomposed fields.
The correlators in SU(2), U$_1$ and $\theta_\mu (n)^{Ds}$ shows
similar exponential decay and existence of light pion mode.
On the other hand, in the regular part, the correlator behaves as a product
of free staggered fermion. In addition to previous results
on chiral condensate at finite temperature \cite{Miya},
these results give support for an
idea that monopole condensation is also responsible for chiral symmetry
breaking.
We find an evidence of strong correlation between instantons
and monopoles. The field dominated by monopoles carrys topological charge
similar to the original SU(2) field . On the other hand ,
The field dominated by photons shows no topological charge.
Present results suggest possibility that
suitable effective theory of monopole condensation
can describe both chiral symmetry breaking of QCD as well as confinement.
\cite{Yotu}
\topskip 0cm
\section{Acknowledgements}
The authors acknowledge T.Suzuki, H.Toki, H.Suganuma, S.Hioki and S.Kitahara
for discussions and comments.
\section{References}
|
\section{#1}}
\def\firstsubsec#1{\obsolete\firstsubsec \subsection{#1}}
\def\thispage#1{\obsolete\thispage \global\pagenumber=#1\frontpagefalse}
\def\thischapter#1{\obsolete\thischapter \global\chapternumber=#1}
\def\nextequation#1{\obsolete\nextequation \global\equanumber=#1
\ifnum\the\equanumber>0 \global\advance\equanumber by 1 \fi}
\def\afterassigment\B@XITEM\setbox0={\afterassigment\B@XITEM\setbox0=}
\def\B@XITEM{\par\hangindent\wd0 \noindent\box0 }
\def\catcode`\^^M=5{\catcode`\^^M=5}
\catcode`@=12
\message{ by V.K.}
\everyjob{\input myphyx }
\chapter{Introduction}
Gravitational lensing has been known for three quarters of a century
and has been extensively used in studies of quasars and clusters of
galaxies. In these cases, the lenses are compound lenses consisting of
extended objects such as galaxies and clusters of galaxies. The recent
gravitational microlensing experiments looking for baryonic dark matter
(Alcock, \etal, 1993, Aubourg, \etal, 1993, Udalski, \etal, 1993)
study nearby stars ($\mathrel{\mathpalette\@versim<} 50 $kpc) as source stars, however,
and the lenses are point lenses.
They are mostly single lenses but some of them have turned out
to be binary lenses (Udalski, \etal, 1994, Bennett, \etal, 1995,
Alcock, \etal, 1995, Alard, \etal, 1995).
The detection rate of binary lenses is expected
to be about $10\%$ according to Mao and Paczy{\'n}ski\ (1991).
One interesting class of binary
lenses is that of planetary systems, and
there is an effort to look for extra solar earth mass planets using
gravitational microlensing (Tytler \etal, 1995).
Obviously, the observational advantage of having to deal with point
lens systems is that the simplicity allows high precision experiments.
From a theorist's point of view, the current and proposed microlensing
experiments have brought a necessity to study fine details of the
cleanest lens systems, which in turn would lead to a more
sophisticated understanding of more uncertain lens systems.
It is well known that when two masses are nearby, they produce a
gravitational lens with a rich structure that depends on the
mass ratio and separation between them and the lens structure is
best represented by the caustics. The caustic curves are closed cuspy
loops, and neither do they self-intersect nor nest each other.
\FIG\ternary{The source plane of a symmetric ternary lens with the
seperation (distance between two masses) $l = 1$: The caustics
divide the source plane into a heirachy of ``nested" domains.
\ The domain marked by $\times$ is the {\it maximal domain} ${\cal D}^3$,
and the number of images of a source in the area is 10.
Incidentally, $\times$ is at the center of the lens positions.} %
(If $n\ge3$, the lenses are general enough to feature
self-intersections and nesting of the caustic loops.
See figure \ternary\ for an example of a ternary lens.)
A caustic curve of a binary lens defines an {\it inside}
and an {\it outside} and the number of images of a source is five
{\it inside} a caustic loop and three {\it outside}.
Infinity lies in the {\it outside} domain and the number of the
images of a source at infinity is three, one at the source position
and the other two at the lens points.
Obviously, the source at infinity is
unamplified and the total microlensing amplification $A_{\rm tot} (\infty) =1$.
The amplification of the image at the source position
is one and that of the image at each lens position is zero.
The amplification of a {\it positive image}
($\equiv$ an image with positive parity)
is always no less than 1, and the total amplification $A_{\rm tot} \ge 1$
in the {\it outside} domain. In an {\it inside} domain, a source produces
two positive images and hence $A_{\rm tot} > 2$. In fact,
$A_{\rm tot} \ge 3$ {\it inside} a caustic as shown recently
by Witt and Mao (1995). Their main result was the following relation.
$$
\sum_{\rm images} {1\over J} = 1 \ , \eqn\Jone
$$
where $J$ is the Jacobian determinant of the lens equation.
In order to demonstrate \Jone, the authors converted the binary lens
equation inside a caustic into a fifth order polynomial equation of
$J^{-1}$ using an algebraic computational package {\it Mathematica}.
They also made an empirical suggestion that there is only one
binary lens configuration that accommodates the infimum amplification 3
inside a caustic.
Here we show that there are actually one-parameter family of binary
lenses where $A_{\rm tot} \ge 3$ is saturated {\it inside} a caustic.
We also present a simple algebraic proof of \Jone\ and show that
a similar ``sum rule'' holds in the ``maximal domains" of arbitrary
$n$-point lens systems. This leads to a necessity to look at the source
planes as consisting of {\it graded caustic domains} $\{{\cal D}^m \}$.
In section 4, we discuss the hierarchical structure
of the source plane of an arbitrary $n$-point lens
and infimum microlensing amplifications of
the sources in the ``maximal domains."
\chapter {Inside the Caustics of Binary Lenses}
A lens equation is a mapping from an image position to a source
position on the lens plane transverse to the line of sight.
If we set the distance scale of the lens plane by the Einstein ring
radius of the total mass, the normalized lens equation is given by
(Bourassa, Kantowski and Norton, 1973)
$$
\omega = z - {\epsilon_1\over (\bar z - x_1)}
- {\epsilon_2\over (\bar z - x_2)} \ , \eqn\leqtwo
$$
where $z$ and $\omega$ are the image and source positions in complex
coordinates, $x_1$ and $x_2$ are the lens positions on the real axis,
and $\epsilon_1$ and $\epsilon_2$ \ ($\epsilon_1 + \epsilon_2 =1$) are
the fractional masses of the lens elements.
The Jacobian matrix elements of the lens equation \leqtwo\ are
$$
\partial_z\omega =1 \quad , \quad
\partial_{\bar z}\omega = {\epsilon_1\over(\bar z - x_1)^2}
+ {\epsilon_2\over(\bar z - x_2)^2} \ \equiv \bar\kappa \ ,
\eqn\eqomega
$$
and the Jacobian determinant is
$J = |\partial_z \omega|^2 - |\partial_{\bar z}\omega|^2
= 1 - |\partial_{\bar z}\omega|^2 \le 1 $.
Lensing is an inverse mapping of the lens equation
and the inverse Jacobian matrix elements are given by
$$
\partial_{\omega} z = {1\over J} \quad , \quad
\partial_{\bar\omega} z = - {\bar\kappa \over J} \ .
\eqn\eqomegatwo
$$
For a positive image, $ 0\le J \le 1$ and the amplification
$A = |J|^{-1} \ge 1$.
A source inside a caustic has two positive images, and the total
amplification of the positive images $ A_+ \ge 2$.
Therefore, combined with \Jone, the total amplification $A_{\rm tot} \ge 3$.
In order to find all the lens parameters where the inequality
$A_{\rm tot} \ge 3$ is saturated, we note that
the minimum $A_{\rm tot} = 3$ can occur only if both of the positive
images have $J = 1$ (or $\partial_{\bar z}\omega = 0$).
They are the two finite limit points of a binary lens (Rhie, 1995a).
If $l = |x_1 - x_2|$ denotes the separation between the lens masses,
$$
z_{\ast\pm}
= x_{\rm acm} \pm i \sqrt{\epsilon_1 \epsilon_2}\,l \ ;
\quad x_{\rm acm} \equiv \epsilon_1 x_2 + \epsilon_2 x_1 \ .
\eqn\bstar
$$
The limit points $z_{\ast\pm}$ are the images of $\omega_{\ast\pm}$
which we can calculate from the lens equation \leqtwo. If we let
$\epsilon\equiv \epsilon_1 - \epsilon_2$,
$$
\omega_{\ast\pm} = x_{\rm acm}
- {\epsilon \over l^2}(x_2-x_1)
\pm i \sqrt{\epsilon_1\epsilon_2}\,\left(l - {2\over l}\right) \ ,
\eqn\eqlimits
$$
and $\omega_{\ast+}$ and $\omega_{\ast-}$ have to be degenerate
in order to be one source producing two images at the limit points.
$$
\omega_{\ast+} = \omega_{\ast-} \quad \Rightarrow \qquad
l = \sqrt{2} \quad ; \qquad
\omega_{\ast} = {1\over 2} ( x_1 + x_2) \ . \eqn\found
$$
In other words, a family of binary lenses with seperation $l =\sqrt{2}$
accommodate a source position with the infimum amplification 3 inside
a caustic
and the source position is at the mid-point of the lens positions
irrespectively of the mass ratio that parameterizes the family.
The fact that there are two positive images implies that the source
in \found\ is inside a caustic. (In a binary lensing, the number of
negative images is always bigger than that of positive images by one.)
The three negative images are on the lens axis and all have $J=-3$.
If the source is at the origin ($x_1 + x_2 = 0$), the lens
equation for the negative images (on the real axis) is given by
$$
0 = x (x^2 - {3\over 2}) + {\epsilon\over\sqrt{2}}
\ , \eqn\eqone
$$
where $x_2 > x_1$\ is assumed without loss of generality.
The Jacobian determinant is given by
$$
J = 1 - \left({(x - x_{\rm acm})^2
+ \epsilon_1\epsilon_2 l^2\over (x-x_1)^2 (x-x_2)^2} \right)^2
\equiv 1 - |\kappa|^2 \ . \eqn\eqJ
$$
When the lens is symmetric ($\epsilon = 0$), the images are at
$ x = 0 \ , \ \pm\sqrt{3/2}$ and it is easy to verify
that $J = -3$ for each image. For an arbitrary $\epsilon$,
one can reduce the numerator and denominator of $\,|\kappa|^2\,$ into
second order polynomials using \eqone\ to find that $|\kappa|^2 = 4$
nd so $J = -3$.
The family of lenses defined by $l=\sqrt{2}$ have caustics
consisting of either one connected loop or two. (For a three-loop
caustic, $l <1$. See Rhie, 1995a).
Two-loop caustics occur when the minor mass is small,
and the source position for the infimum 3 is in the caustic near the minor
mass. Inside the caustic (very) near the dominant mass, the amplification
goes to infinity as the mass asymmetry grows because the
caustic approaches that of a single lens.
Here we demonstrate that \Jone\ does not hold when the source is
{\it outside} -- the non-maximal domain of a binary lens -- except when
the source is at infinity. Let's consider a source located
at the mid-point of a symmetric lens.
When $l < 2$, the source lies inside the caustic, and there are five
images: two off the axis ($z=\pm\sqrt{d^2-1}$; $d = l/2$),
and the other three ($z=0, \pm\sqrt{d^2+1}$) on the axis.
As the separation $l$ grows beyond the bifurcation condition
$l=2$, the caustic splits into two
closed loops and the source falls in the {\it outside} domain. Thus,
two out of the five images should disappear, and they are the
two solutions $z=\pm\sqrt{d^2 -1}$, which become real at $l = 2$.
Both have $J = - 4d^2(d^2-1)\ ( < 0 {\rm for} l > 2 )$. Therefore,
the other three images have $\sum_{\rm images} J^{-1} > 1$
($ \rightarrow 1$ as $d \rightarrow \infty$). Of course, the total
amplification of the three images of a source {\it outside}
can be dominated by negative images and $\sum_{\rm images} J^{-1} < 1$.
An example is a source near the off-axis cusps which appear when
the caustic bifurcates into three loops.
\chapter {$\sum J^{-1} = 1$ for the Maximum Number of Images}
The lens equation \leqtwo\ is a mapping from an image position to a
source position: $z, \bar z \mapsto \omega$, and lensing is the inverse
of the lens mapping: $\omega, \bar\omega \mapsto z$ which is
either triple-valued or quintuple-valued.
If we let $z_1 \equiv z - x_1$, $z_2 \equiv z - x_2$ and
$z_{\circ} \equiv z - x_{\rm acm}$, the binary lens equation looks
quite simple.
$$
\omega = z - {\bar z_{\circ}\over \bar z_1 \bar z_2}
\equiv z - f(\bar z) \ . \eqn\leqthree
$$
We note that the effect of the masses $\epsilon_j: j=1,2$\ is coupled
only to the conjugate variable $\bar z$ and the relation between $\omega$
and $z$ is universal. We will see below that
this universality is the underlying reason for the constraint \Jone,
and thus, it holds not only for binary lenses but also for arbitrary
$n$-point lenses.
In order to find the inverse function $z(\omega,\bar\omega)$,
we replace $\bar z$ in \leqthree\ by $\bar\omega + f(z)$:
$ \omega = z - f(\bar\omega + f(z))$ to find that $z$ satisfies
a fifth order polynomial
equation, $0 = g(z;\omega, \bar\omega)$\ (Witt, 1990).
If $\omega_j; \ j = {\circ}, 1, 2$ is defined similarly to $z_j$,
$$
g(z;\omega,\bar\omega) = (z-\omega)(z_{\circ}+\bar\omega_1 z_1 z_2)
(z_{\circ}+\bar\omega_2 z_1 z_2)
- z_1z_2(z_{\circ}+\bar\omega_{\circ}z_1z_2) \ .
\eqn\eqz
$$
$g=0$ is an analytic equation in $z$ and always has five solutions.
Thus, $g=0$ is not equivalent to
the original lens equation when the source is {\it outside} where
there are only three images.
However, in the {\it inside} domains we are interested in here,
the sources result in the maximum number of images 5
and so we can consider $0=g$ as the lens equation.
If $a_j (\omega, \bar\omega); j =1,5$ are the zeros of $g$,
we can factorize $g$ as
$ g (z;\omega,\bar\omega) = a_{\circ}\Pi_{j=1}^5 (z-a_j)$,
where $a_{\circ} = \bar\omega_1 \bar\omega_2$. \
Now, we note from \eqomegatwo\ that $J^{-1}$ is a very primary
quantity we can calculate easily. The value of $J^{-1}$ -- the amplification
combined with the parity -- at an image position $a_j$ is
$$
J^{-1} (a_j) = \partial_{\omega} z (a_j) =
- {\partial_\omega g\over\partial_z g}(a_j)
= \partial_\omega a_j (\omega, \bar\omega) \ .
\eqn\eqJinv
$$
By summing it over all the images using \eqz, we find the ``sum rule"
completing the proof.
$$
\sum_{j=1}^5 J^{-1} (a_j)
= \partial_\omega \sum_{j=1}^5 a_j(\omega, \bar\omega) = 1 \ .
\eqn\eqsumj
$$
The relevent terms in $g$ for the second equality are the highest order
term for the whole polynomial and the next highest order term depending on
$\omega$: $(z-\omega)z^4 \bar\omega_1\bar\omega_2$. The constant 1 on the RHS
is determined by the relation between $\omega$ and $z$ in the factor
$(z - \omega)$.
For a single lens, there are two images and $g$ is a quadratic polynomial
in $z$. Since $g=0$ is saturated by the images (or $g=0$ is equivalent
to the lens equation), \eqsumj\ holds for a single lens. Or, one can
satisfy oneself by solving the quadratic lens equation and adding up
$J^{-1}$ for the solutions. ($J = 1-|z|^{-4}$, where the lens is at $z =0$.)
Now, the generalization of \eqsumj\
for an arbitrary class of $n$-point lens systems
is only a matter of bookkeeping.
The lens equation is given by
$\omega = z - \sum_{j=1}^n \epsilon_j/\bar z_j \ ; \ \
\sum_{j=1}^n \epsilon_j = 1 \ . $
We note that $f(\bar z)$ in \leqthree\ is a quotient
of a first order polynomial by a second order polynomial:
$f = f_1/f_2$. For an $n$-point lens, $f(\bar z)$ is a quotient
of an $(n-1)$-th order polynomial by an $n$-th order polynomial:
$f = f_{n-1}/f_{n}$. For a binary, the polynomial in \eqz\ \
$g(z;\omega, \bar\omega) \equiv g_2 \
\supset (z-\omega) \Pi_{j=1}^2(f_1 + \bar\omega_j f_2) \ \supset
a_{\circ} (z-\omega) z^4$.
For an $n$-point lens, the corresponding polynomial
$g(z;\omega, \bar\omega) \equiv g_n
\ \supset (z-\omega) \Pi_{j=1}^n(f_{n-1} + \bar\omega_j f_n)
\ \supset a_{\circ} (z-\omega) z^{n^2}$.
Therefore,
$$
\sum_j J^{-1}(a_j)
= \partial_\omega \sum_j a_j (\omega, \bar\omega) = 1 \ . \eqn\eqsumjn
$$
We emphasize that the second equality of \eqsumjn\ holds only when
the summation is over all the $n^2+1$ images, \hbox{\it i.e.}} \def\etc{\hbox{\it etc.}, in the ``maximal
domains." Only in the ``maximal domains", the lens equation is equivalent
to $g=0$.
\chapter{Graded Caustic Domains}
There is no ``inside a caustic" in a single lens ($n=1$) because
the caustic is point -- a ``collapsed caustic loop".
For binary lenses ($n=2$),
``inside a caustic" is ``a maximal domain" and the source plane consists
of one, two or three ``islands" of ``maximal domains" surrounded by
the {\it outside}. When $n\ge3$, the caustics show more
complicated structures as we have seen one of whose examples
in figure \ternary.
Now, how do we sort these apparently complicated structures and understand
them intuitively? Actually, it is rather simple to see that a source
plane consists of {\it graded caustic domains} -- namely, {\it domains}
with well-defined {\it degrees}. We only need to
recall a couple of well-known facts. \ \
First, the number of images of a source remains
the same if the source does not cross a caustic curve. So, we can spell out
the self-evident definition of a {\it domain}:
{\it A domain of a source plane
is an area bounded by but devoid of caustic curves}. \ \
Second, the number of images changes by two
-- one positive and one negative --
when a source crosses a caustic curve.
Since caustic curves are oriented as determined by the phase angle of $\kappa$
we encountered in section 2, we can consider the sign of caustic crossings.
Then, given two randomly chosen domains, the algebraic sum of the number
of caustic crossings with sign $\pm$
is independent of the paths connecting the two domains (see Rhie, 1995a for
more details).
Now, define the {\it degree} of a caustic domain:
The {\it outside} domain has {\it degree}
zero and is denoted ${\cal D}^{\circ}$.
{\it If the algebraic sum of the caustic crossing between a {\it domain}
and {\it outside} is $m$, the {\it domain} has {\it degree} $m$
and is denoted ${\cal D}^m$.} (Lensing is a short range phenomenon,
and ${\cal D}^{\circ}$ can consist of many domains where only one of them
includes infinity.) Some of the corollaries are : \ \
1)\ ${\cal D}^{m}$ is always nested in ${\cal D}^{m-1}$'s
and the number of images of a source in ${\cal D}^m$ is $n+1+2m$.
\ \ 2)\ A {\it maximal domain} is where a source has $n^2+1$ images
by definition. Now, the number of images of a source
in ${\cal D}^{\circ}$ is $n+1$
and thus the {\it degree} of a {\it maximal domain} is $n(n-1)\over 2$.
If the maximum {\it degree} of the domains of an $n$-point lens is $N(n)$,
the source plane is \ $\cup_{m=\circ}^{N(n)} \{{\cal D}^m\}$ where
$N(n) \le n(n-1)/2$. We write ${\rm sup}N(n) = n(n-1)/2$.
\ Let's look at the example in figure \ternary:
The source plane is made of one ${\cal D}^{\circ}$, nine
${\cal D}^1$'s, one ${\cal D}^2$ and one ${\cal D}^3$.
As the seperation increases, the triple lens should converge to
three single lenses: At a large separation, the source plane has three
${\cal D}^1$'s near the lens positions. The point we are getting at is
that the fact that ${\cal D}^1$'s in the ``ocean" of ${\cal D}^{\circ}$ is
a generic feature of the source plane of an arbitrary $n$-point lens
when the ``correlations" of the lens masses is weak.
As the masses come close and
their ``correlations" become strong, the ${\cal D}^1$'s merge and generate
higher hierarchical structures.
In the case of binaries, however, ${\cal D}^1$ is also {\it maximal} and
merging does not result in higher degrees.
There is only one positive image for a source at infinity and thus for
a source anywhere in ${\cal D}^{\circ}$.
Therefore, the number of positive images of a source in ${\cal D}^m$
is $1+m$, and the total amplification of the positive images is
$ A_+ ({\cal D}^m) \ge 1 + m$.
A universal constraint on negative images is that they have
non-vanishing amplifications unless the source is at infinity
$\in {\cal D}^{\circ}$.
Therefore, $ A_{\rm tot}({\cal D}^m) > 1 + m$ when $m \ge 1$
(or ``inside the caustics").
Inside a `maximal domain', the constraint \eqsumjn\ applies,
and $ A_{\rm tot} \ge n(n-1)+1$. Whether this inequality is saturated is
another question, and we will discuss this briefly below for $n\ge3$.
$ A_{\rm tot} = n(n-1)+1$ only if all the positive images are at the limit
points. Since there are $2(n-1)$ (finite) limit points,
$n$ has to satisfy that $2(n-1)=1+n(n-1)/2$.
In other words, $n=3$ is the only possibility besides the known cases of
$n = 1, 2$. Indeed, we find that a source at the center of a symmetric
ternary lens with
separation (distance between two masses)
$l = {\root 6 \of 2} \approx 1.12246$ has four
positive images at the (finite) limit points. Therefore, $A_+ = 4$
and $A_{\rm tot} = 7$. It should be an
interesting question whether the solution we have found is unique or
one of many. For $n \ge 4$, the total amplification
of the maximum number of images is formally $A_{\rm tot} > n^2+1-n $.
It should be interesting to investigate the significance
of the high degree domains in terms of the size and frequency of
the appearance in the source planes of multiple point lens systems.
Since this paper was submitted, the MACHO experiment has detected a binary
event toward the LMC (Bennett \etal, 1996, astro-ph/9606012) that shows
the amplification
between two caustic crossings less than 2. We have mentioned above
that $ A_{\rm tot} > 2$ inside any caustic of any $n$-point lens system.
Therefore, one can {\it definitely} conclude that the event was
``contaminated" by other than geometric effect due to gravity (such as
the third component of the lens).
If the ``contamination" is due to blending of the source star that lies
in the bar of the LMC, it may be verified by the centroid shift
of the source star (MACHO, private communication).
\singl@true\doubl@false\spaces@t
\ack
We thank D. Bennett for valuable input.
This work was supported in part by
the U.S. Department of Energy at the Lawrence Livermore
National Laboratory under contract No. W-7405-Eng-48.
\bigskip
\bigskip
\title {REFERENCES}
\def{\it A. \& A.}{{\it A. \& A.}}
\def{\it ApJ}{{\it ApJ}}
\def{\it Ann. Rev. Astron. Astrophys.}{{\it Ann. Rev. Astron. Astrophys.}}
\def{\it MNRAS}{{\it MNRAS}}
\def{\it AAp}{{\it AAp}}
\def\par\hangindent=1cm\hangafter=1\noindent{\par\hangindent=1cm\hangafter=1\noindent}
\parskip 0pt
\par\hangindent=1cm\hangafter=1\noindent Alard, C., Mao, S., and Guibert, J., 1995, submitted to {\it A. \& A.}
\par\hangindent=1cm\hangafter=1\noindent Alcock, C., Akerlof, C.W., Allsman, R.A., Axelrod, T.S., Bennett, D.P.,
Chan, S., Cook, K.H., Freeman, K.C., Griest, K., Marshall, S.L., Park, H.-S.,
Perlmutter, S., Peterson, B.A., Pratt, M.R., Quinn, P.J., Rodgers, A.W.,
Stubbs, C.W., and Sutherland, W., 1993, {\it Nature}, {\bf 365}, 621.
\par\hangindent=1cm\hangafter=1\noindent Alcock, C., Allsman, R.A., Axelrod, T.S., Bennett, D.P.,
Chan, S., Cook, K.H., Freeman, K.C., Griest, K., Marshall, S.L.,
Perlmutter, S., Peterson, B.A., Pratt, M.R., Quinn, P.J., Rodgers, A.W.,
Stubbs, C.W., and Sutherland, W., 1995, {\it ApJ}, in press.
\par\hangindent=1cm\hangafter=1\noindent Aubourg, E., Bareyre, P., Brehin, S., Gros, M., Lachieze-Rey, M.,
Laurent, B., Lesquoy, E., Magneville, C., Milsztajn, A., Moscosco, L.,
Queinnec, F., Rich, J., Spiro, M., Vigroux, L., Zylberajch, S., Ansari, R.,
Cavalier, F., Moniez, M., Beaulieu, J.-P., Ferlet, R., Grison, Ph.,
Vidal-Madjar, A., Guibert, J., Moreau, O., Tajahmady, F., Maurice, E.,
Prevot, L., and Gry, C., 1993, {\it Nature}, {\bf 365}, 623.
\par\hangindent=1cm\hangafter=1\noindent Bennett, D.P. \etal, AIP Conference Proceedings {\bf 336}:
Dark Matter, eds., S. S. Holt and C. L. Bennett, 1995, p.77.
\par\hangindent=1cm\hangafter=1\noindent Bourassa, R. R., Kantowski, R, Norton, T. D., 1973,
{\it ApJ}, {\bf 185}, 747.
\par\hangindent=1cm\hangafter=1\noindent Mao, S. and Paczy{\'n}ski, B., 1991, {\it ApJ}, {\bf 374}, L37.
\par\hangindent=1cm\hangafter=1\noindent Rhie, S., 1995a, preprint.
\par\hangindent=1cm\hangafter=1\noindent Rhie, S., 1995b, preprint.
\par\hangindent=1cm\hangafter=1\noindent Tytler, D., \etal, 1995, in preparation.
\par\hangindent=1cm\hangafter=1\noindent Udalski, A., Szymanski, M., Kaluzny, J., Kubiak, M., Krzeminski, W.,
Mateo, M., Preston, G.W., and Paczy{\'n}ski, B., 1993,
{\sl Acta Astronomica} {\bf 43}, 289.
\par\hangindent=1cm\hangafter=1\noindent Udalski, A., Szymanski, M., Kaluzny, J., Kubiak, M., Krzeminski, W.,
Mateo, M., Preston, G.W., and Paczy{\'n}ski, B., 1994, {\it Acta Astronomica},
{\bf 44}, 165.
\par\hangindent=1cm\hangafter=1\noindent Witt, H.J., 1990, {\it A. \& A.}, {\bf 236}, 311.
\par\hangindent=1cm\hangafter=1\noindent Witt, H. J. and Mao, S., 1995, {\it ApJ}, {\bf 447}, L105.
\vfil\break
\figout
\vfil\break
|
\section{Introduction}
Weak decays of hypernuclei provide a laboratory for
investigating kaon-nucleon interactions.
In free space, a hyperon such as $\Lambda$, $\Sigma$ or $\Xi$
will decay predominantly through
a mesonic mode, e.g. $\Lambda\rightarrow N\pi$.
However, when the hyperon is bound inside a nucleus with $A \sim
12$ or
larger,
this decay mode is suppressed by Pauli-blocking of the final
state
nucleon (produced with $\vec{p}_N \sim 100 $ MeV/c, much less
than
the Fermi-momentum $\vec{p}_F \sim 280$ MeV/c).
A competing process that does not suffer significantly
from Pauli-blocking is nonmesonic weak scattering, e.g.
$\Lambda N\rightarrow NN$, in which the final state nucleons have
$\vec{p}_N \sim 400$ MeV/c.
These processes occur at low momentum scales where
QCD is in the nonperturbative regime, so the
structure and decays of $\Lambda$, $\Sigma$, and $\Xi$
hypernuclei have been investigated using various phenomenological
models,
e.g.
\cite{MG84,D86,OS8586,CHK83,HK86,N88,MDG88,C90,RMBJ92,D92,ITO94,MS94,MJ95}.
Available experimental observables for analyzing such systems
include spin-averaged decay rates,
the proton asymmetry from polarised hypernuclei,
and the ratio of neutron induced to proton induced
decay widths.
The later proves especially difficult to describe using
hadronic models, implying that we do not yet have a complete
understanding of the dynamics of these systems.
The weak scattering process receives
contributions from long distance meson exchange diagrams
and from short distance (compared to the scale of chiral symmetry
breaking,
$\Lambda_\chi \sim$ 1 GeV) four-baryon contact terms.
The leading meson exchange graphs are the one-pion
exchange (OPE) followed by
one-kaon exchange (OKE), one-eta exchange (OEE)
and two-pion exchange (TPE) and so forth.
The OPE amplitudes can be determined relatively
well because both the weak and strong vertices have been
experimentally determined.
The OKE and OEE graphs are expected to be the next largest
contribution and
model computations\cite{RMBJ92}\ show that a significant
contribution ($\sim 30\%$) to the
nonmesonic decay mode may come from the OKE amplitude.
More importantly, it has been demonstrated that
the ratio of neutron-induced to proton-induced
decay widths of $\Lambda$ hypernuclei is sensitive to
the weak OKE amplitude \cite {benntalk95},
which, for the currently used values for the NNK vertices,
significantly cancels the contribution of the OPE amplitude.
The resulting small ratio found for $^{12}_\Lambda C$ \cite {benntalk95}
is not consistent with what is seen experimentally
\cite{BNL91,KEK92,Ejiri95},
although experimental uncertainties are large.
It also appears that vector meson exchange (e.g. $K^*, \rho, ...$)
contributes to this ratio \cite {benntalk95}.
Such exchanges would be included in local four-baryon
$\Delta s=1$ operators in chiral perturbation theory.
It is not possible to make a direct experimental determination of
the NNK weak couplings
that appear in the nonmesonic decay amplitudes; instead, flavour
SU(3) is used
to relate these couplings to the weak pion couplings.
Previous analyses of hypernuclear decay have either ignored SU(3)
breaking
or assigned an arbitrary $30\%$ uncertainty
to these couplings as an estimate of the SU(3) breaking, eg.
\cite{MS94} .
We will estimate the size of SU(3) breaking in the NNK amplitudes
using
chiral perturbation theory.
Understanding the NNK weak interactions may shed light on a
troubling
situation encountered in the decay of free hyperons
outside the role they play in hypernuclear decay.
Both the S-wave and P-wave amplitudes for $\Delta s=1$ hyperon
decay
are well studied experimentally.
The S-wave amplitudes are adequately described, even at
tree level, by a weak
operator transforming as an $(8_L,1_R)$ under chiral
${\rm SU(3)_L}\otimes {\rm SU(3)_R}$.
A long standing problem is that the P-wave amplitudes are not
well
reproduced
at tree level using the coupling constants extracted from the
S-wave
amplitudes. A one-loop calculation of the leading SU(3)
corrections to
hyperon decay,
performed in ref.\cite{ej}, showed that this situation is not
improved by
including the leading
terms nonanalytic in the strange
quark mass. Further, these corrections
change the tree level prediction
of P-wave amplitudes by $100\%$ (typical SU(3) breaking
corrections are $\sim 30\%$), causing concern \cite{BSW85a,georgi}
that chiral perturbation theory may not be valid for
such processes.
It was suggested in ref. \cite{ej} that the problem may instead
be that the
weak coupling constants extracted from S-wave fits lead to
(accidental)
cancellations between the tree level P-wave amplitudes.
Large SU(3) breaking effects are then a result of small
tree level
amplitudes,
and not a breakdown of chiral perturbation theory.
Since there is
only one graph that contributes to P-wave NNK interactions at
tree level, such accidental cancellations are absent.
Experimental determination of these weak NNK vertices would
provide
an indication of the applicability of chiral perturbation theory
to such processes.
\section{The Chiral Lagrangian for Nonleptonic Interactions}
At the momentum transfers characteristic of nonmesonic
hypernuclear decay, $p < \Lambda_\chi$, the relevant degrees of
freedom
are the lowest mass octet and decuplet baryons and the
pseudo-Goldstone bosons
$\pi$, $K$, and $\eta$.
The low energy strong interaction of the these hadrons
is described by the Lagrange density
\begin{eqnarray}\label{strongl}
{\cal L}^{st} &=& i {\rm Tr} \bar B_v \left(v\cdot {\cal D}
\right)B_v
+ 2 D\ {\rm Tr} \bar B_v S_v^\mu \{ A_\mu, B_v \}
+ 2 F\ {\rm Tr} \bar B_v S_v^\mu [A_\mu, B_v]
\nonumber \\
&&- i \bar
T_v^{\mu} (v \cdot {\cal D}) \ T_{v \mu}
+ \Delta m \bar T_v^{\mu} T_{v \mu}
+ {\cal C} \left(\bar T_v^{\mu} A_{\mu} B_v + \bar B_v A_{\mu}
T_v^{\mu}\right)
\nonumber\\
&& + 2 {\cal H}\ \bar T_v^{\mu} S_{v \nu} A^{\nu} T_{v \mu}
+ {f^2 \over 8} {\rm Tr} \partial_\mu \Sigma \partial^\mu
\Sigma^\dagger
+ \mu {\rm Tr} \left( m_q\Sigma + m_q^\dagger\Sigma^\dagger \right) \
+\ \cdots \ \ \ \ ,
\end{eqnarray}
where $f$ is the meson decay constant, $m_q$ is the light quark
mass matrix,
${\cal D_\mu}= \partial_\mu+[V_\mu, \; ]$
is the covariant chiral derivative
and use is made of the vector and axial vector chiral currents
\begin{eqnarray}
V_\mu&=&{1 \over 2} (\xi\partial_\mu\xi^\dagger +
\xi^\dagger\partial_\mu\xi)
\nonumber \\
A_\mu&=&{i \over 2} (\xi\partial_\mu\xi^\dagger -
\xi^\dagger\partial_\mu\xi)
\ \ \ .
\end{eqnarray}
The dots in Eq.~\ref{strongl}\ represent higher dimension operators
(involving
more derivatives and insertions of the light quark mass matrix)
whose contributions are suppressed by inverse powers of
$\Lambda_\chi$.
The octet baryon field of four-velocity $v$ is denoted by $B_v$
and has SU(3)
elements
\begin{eqnarray}\label{octet}
B_v =
\pmatrix{ {1\over\sqrt2}\Sigma_v^0 + {1\over\sqrt6}\Lambda_v &
\Sigma_v^+ & p_v\cr \Sigma_v^-& -{1\over\sqrt2}\Sigma_v^0 +
{1\over\sqrt6}\Lambda_v&n_v\cr \Xi_v^- &\Xi_v^0 &-
{2\over\sqrt6}\Lambda_v
\cr }
\ \ \ ,
\end{eqnarray}
and the decuplet baryons appear as the elements of the (totally
symmetric) $T_v$:
\begin{eqnarray}\label{decuplet}
& & T^{111}_v = \Delta^{++}_v, \ \ T^{112}_v =
{1\over\sqrt{3}}\Delta^{+}_v,
\ \ T^{122}_v = {1\over\sqrt{3}}\Delta^{0}_v, \ \ T^{222}_v =
\Delta^{-}_v,
\nonumber\\
& & T^{113}_v = {1\over \sqrt{3}}\Sigma^{*+}_v,\ \
T^{123}_v = {1\over\sqrt{6}}\Sigma^{*0}_v, \ \
T^{223}_v = {1\over\sqrt{3}}\Sigma^{*-}_v,
\ \ T^{133}_v = {1\over\sqrt{3}}\Xi^{*0}_v,
\nonumber\\
& & T^{233}_v = {1\over\sqrt{3}}\Xi^{*-}_v, \ \ T^{333}_v =
\Omega^-_v \ \ \ .
\end{eqnarray}
The octet of pseudoscalar pseudo-Goldstone bosons
resulting from the spontaneous breaking of chiral symmetry appear
in
the $\Sigma$ field, with
\begin{eqnarray}
\Sigma = \xi^2= {\rm exp}\left( {2 i M\over f} \right) \ \ \ ,
\end{eqnarray}
where
\begin{eqnarray}
M =
\left(\matrix{{1\over\sqrt{6}}\eta+{1\over\sqrt{2}}\pi^0&\pi^+&K^
+\cr
\pi^-&{1\over\sqrt{6}}\eta-{1\over\sqrt{2}}\pi^0&K^0\cr
K^-&\overline{K}^0&-{2\over\sqrt{6}}\eta\cr}\right)
\ \ \ \ .
\end{eqnarray}
The strong couplings constants $F, D, {\cal C}$ and ${\cal H}$
have
been determined from one-loop computations of axial matrix
elements
between octet baryons \cite{mj}
and strong decays of decuplet baryons \cite{bss}.
The $\Delta s=1$ weak interactions of the pseudo-Goldstone bosons
and the lowest lying baryons are described, assuming octet
dominance, by the Lagrange density
\begin{eqnarray}\label{weakl}
{\cal L}^{\Delta s=1}_v &=&
G_Fm_\pi^2 f_\pi \Big( h_D {\rm Tr} {\overline B}_v
\lbrace \xi^\dagger h\xi \, , B_v \rbrace \;
+ \;
h_F {\rm Tr} {\overline B}_v
{[\xi^\dagger h\xi \, , B_v ]} \; \nonumber \\ &&
+ h_C {\overline T}^\mu_v
(\xi^\dagger h\xi) T_{v \mu} \; + \;
{ h_\pi \over 8} {\rm Tr} \left( h \, \partial_\mu
\Sigma
\partial^\mu
\Sigma^\dagger \right)
\ + \ \cdots\ \ \ \Big) \ \ \ \ ,
\end{eqnarray}
where
\begin{eqnarray}
h = \left(\matrix{0&0&0\cr 0&0&1\cr 0&0&0}\right) \ \ \ ,
\end{eqnarray}
and the constants $f_\pi$, $h_D, h_F, h_\pi$ and $h_C$ are
determined experimentally.
The pion decay constant is known to be $f_\pi \sim 132$ MeV.
We have inserted factors of $ G_Fm_\pi^2 f_\pi $ in Eq.~\ref{weakl}\
so that the constants $h_D, h_F$, and $h_C$ are
dimensionless and
of order unity.
At tree level, the weak decay of the octet baryons
gives\cite{mj,ej}
$h_D = -0.58$ and $h_F = +1.40$, while the weak
decay of the $\Omega^-$ gives
$h_C \sim 1.4$.
The weak meson coupling $h_\pi$ is determine from nonleptonic
kaon decays
to be $h_\pi = 1.4$ MeV.
The dots denote higher dimension operators
involving more derivatives and
insertions of the light quark mass matrix.
We will determine the S-wave and P-wave amplitudes for weak
NNK interactions, including the SU(3) violating one-loop
corrections.
In the spirit of chiral perturbation theory we compute the
leading
nonanalytic corrections dependent upon the mass of the strange
quark,
of the form $m_s\ln m_s$.
This requires the computation of one-loop graphs involving
kaons, pions, and etas with
octet and/or decuplet baryons. Such graphs are divergent and
regularized in
$n$-dimensions with modified minimal subtraction,
$\overline{MS}$.
The divergences are absorbed by higher dimension operators whose
coefficients
depend upon the renormalization scale.
The sum of the counterterm and the loop graph is scale
independent.
By choosing to renormalize at the chiral symmetry breaking scale,
and using the fact that the
coefficients of the higher dimension operators are analytic
functions of the
light
quarks masses,
the size of these coefficients can be {\it estimated} using naive
dimensional
analysis.
In the chiral limit the contributions from the higher dimension
operators are
subdominant compared to the logarithms that arise from the loop
graphs
involving the lowest dimension operators.
It is these chiral logarithms that we compute in this work.
For physical values of the kaon, pion, and eta masses, these
quantities
represent only
an estimate of the size of
SU(3) breaking effects; contributions from local counterterms
will be
of the same order. Unfortunately, the coefficients of the
counterterms are not directly computable from the
chiral Lagrangian and must be determined experimentally.
The amplitude for the weak $\Delta s=1$ NNK interactions has the
form
\begin{eqnarray}\label{spamp}
{\cal A} = i G_F m_\pi^2\ {f_\pi \over f_K}\ \overline{N}_v \ \left[
{\cal A}^{(S)} +
2 {k \cdot S_v \over \Lambda_\chi} {\cal A}^{(P)} \right] N_v
\end{eqnarray}
where $N_v$ contains the nucleon doublet
\begin{eqnarray}
N_v = \left(\matrix{ p_v\cr n_v }\right) \, \, \, ,
\end{eqnarray}
and $k$ is the outgoing momentum of the kaon.
The amplitudes $ {\cal A}^{(S)}$ and ${\cal A}^{(P)} $ are the
S-wave and
P-wave amplitudes
respectively, and are computed below.
We have chosen to normalize ${\cal A}^{(P)} $ to $\Lambda_\chi$
so that it is
dimensionless.
An explicit factor of $f_\pi/f_K$ appears in Eq.~\ref{spamp}\ and further
we will
distinguish $f_K$ from $f_\pi$ in the expressions arising from the
one loop amplitudes.
There is evidence from other loop computations that this SU(3)
breaking difference
should be included explicitly \cite{jlms93}.
\section{Computation of Amplitudes}
There are three vertices that occur in $\Delta s=1$ weak
nonleptonic
interactions involving nucleons and kaons:
$p \bar p K^0$, $n \bar p K^+$, and $n \bar n K^0$.
They are not independent in the limit of isospin symmetry
and are related by
\begin{eqnarray}
{\cal A}^{(L)}(nnK) - {\cal A}^{(L)}(ppK) = {\cal A}^{(L)}(npK)
\ \ \ ,
\end{eqnarray}
where $L=0$ (S-wave) or $L=1$ (P-wave);
this relation is true for both S-wave and P-wave amplitudes
independently.
The amplitudes
\begin{eqnarray}
{\cal A}^{(L)} = {\cal A}_0^{(L)} + {\cal A}_1^{(L)} + \cdots \ ,
\end{eqnarray}
where the subscript denotes the order in chiral perturbation
theory and the
dots indicate contributions arising from more insertions of the
light quark mass matrix or involving more derivatives.
\subsection{S-Wave Amplitudes}
At tree level (see Fig.~\ref{tree}) the S-wave amplitudes
appear directly from the
first and second
terms in the
weak Lagrangian of Eq.~\ref{weakl}\ ,
\begin{eqnarray}\label{stree}
{\cal A}_0^{(S)} (p \bar p K^0) &=& h_F-h_D \nonumber \\
{\cal A}_0^{(S)} (p \bar n K^+) &=& h_F+h_D \nonumber \\
{\cal A}_0^{(S)} (n \bar n K^0) &=& 2 h_F
\ \ \ \ .
\end{eqnarray}
Experimental measurements of
hyperon decays are used to find the parameters $h_D$ and $h_F$.
Because the above amplitudes are tree level
expressions, it is appropriate to extract these parameters
using tree level predictions of the chiral
Lagrangian. This gives $h_D$ = --.58 and
$h_F$ = 1.40\cite{ej}. The tree level S-wave amplitudes
are then 2.0, 0.8, and 2.8, respectively.
The numbers will be modified by SU(3) breaking effects, as computed
below.
Direct computation of the loop graphs shown in
Fig.~\ref{swaves} lead to
S-wave amplitudes
\begin{eqnarray}
{\cal A}_1^{(S)} (p \bar p K^0) &=&
{m_K^2 \over 16\pi^2 f_K^2}\ln\left({m_K^2 \over
\Lambda_\chi^2}\right)
\Big( {h_D\over 3} ( 1+13D^2-18DF-27F^2) \nonumber \\
& & \phantom{{m_K^2 \over 16\pi^2 f_K^2}\ln({m_K^2 \over
\Lambda_\chi^2}) +
}
+ {h_F\over 3} ( -7-7D^2+6DF+9F^2) \Big) \nonumber \\
&& + {2 C^2 \over 9}h_C{\cal J}(\Delta m^{\Sigma^*}_N)
- {\cal A}_0^{(S)} (p \bar p K^0) {\cal Z}_\Psi
\ \ \ \ ,
\end{eqnarray}
\begin{eqnarray}
{\cal A}_1^{(S)} (p \bar n K^+) &=&
{m_K^2 \over 16\pi^2 f_K^2}
\ln\left({m_K^2 \over \Lambda_\chi^2}\right)
\Big( {h_D\over 3}(2-4D^2-24F+36F^2) \nonumber \\
& & \phantom{{m_K^2 \over 16\pi^2 f_K^2}\ln({m_K^2 \over
\Lambda_\chi^2}) +
}
+ {h_F\over 3}(2+20D^2-48DF+36F^2) \Big) \nonumber \\
&& - {4 C^2 \over 9}h_C{\cal J }(\Delta m^{\Sigma^*}_N)
- {\cal A}_0^{(S)} (p \bar n K^+) {\cal Z}_\Psi
\ \ \ \ ,
\end{eqnarray}
\begin{eqnarray}
{\cal A}_1^{(S)} (n \bar n K^0) &=&
{m_K^2 \over 16\pi^2 f_K^2} \ln\left({m_K^2 \over
\Lambda_\chi^2}\right)
\Big( h_D( 1+3D^2-14DF+3F^2) \nonumber \\
& & \phantom{{m_K^2 \over 16\pi^2 f_K^2}\ln({m_K^2 \over
\Lambda_\chi^2}) +
}
+ {h_F\over 3} (-5 +13D^2 -42DF + 45F^2) \Big)
\nonumber \\
&& - {2 \over 9} C^2h_C{\cal J}(\Delta m^{\Sigma^*}_N)
- {\cal A}_0^{(S)} (n \bar n K^0) {\cal Z}_\Psi
\ \ \ \ ,
\end{eqnarray}
\noindent where the contribution from wavefunction
renormalization is
given by
\begin{eqnarray}
{\cal Z}_\Psi =
{m_K^2 \over 16\pi^2 f_K^2} \ln ({m_K^2 \over \Lambda_\chi^2})
\Big( 15F^2 - 10FD + {17 \over 3}D^2\Big)
+ C^2 {\cal J}(\Delta m^{\Sigma^*}_N)
\ \ \ \ \ ,
\end{eqnarray}
\noindent and the function ${\cal J}$ is
\begin{eqnarray}
{\cal J}(\delta) = {1 \over 16\pi^2 f_K^2}
&&\left[(m_K^2 -2 \delta^2)\ln\left({m_K^2 \over
\Lambda_\chi^2}\right)
+ 2\delta \sqrt{\delta^2-m_K^2 }
\ln \left( { \delta-\sqrt{\delta^2-m_K^2+ i\epsilon} \over
\delta+\sqrt{\delta^2-m_K^2 + i\epsilon}}\right) \right]
\ \ .
\end{eqnarray}
\noindent For S-waves, $\delta=\Delta
m^{\Sigma^*}_N=m_{\Sigma^*}-m_N$.
The kaon decay constant $f_K$ = 1.22 $f_\pi$.
We have not included wavefunction renormalization of the external
meson
field in our KNN amplitude computations since the kaons do not
appear as
asymptotic states in the weak scattering processes under
consideration.
In order to determine the ${\cal A}^{(S)}$ we insert the axial
coupling
constants
$D, F ,{\cal C}$, and ${\cal H}$ extracted by the one-loop
computations
of
\cite{mj,bss}:
\begin{eqnarray}\label{strongFDCH}
D & = & 0.6\pm 0.1\ \ ,\ \ F = 0.4\pm 0.1\ \ \nonumber \\
{\cal C} & = & -1.2\pm 0.1\ \ ,\ \ {\cal H} = -2.0\pm 0.2
\ \ \ \ ,
\end{eqnarray}
and the weak coupling constants determined at one-loop level
\cite{ej}
\begin{eqnarray}\label{weakFDCH}
h_D & = & -0.35 \pm 0.09 \ \ ,\ \ h_F = 0.86\pm 0.05 \nonumber
\\
h_C & = & -0.36\pm 0.65
\ \ \ .
\end{eqnarray}
The uncertainties in these couplings are treated as uncorrelated
for the purpose of determining the uncertainty in the NNK
amplitudes.
We determine the error in the NNK amplitudes by varying the
parameters over their
allowed range and require that the choice of parameters
reproduce the $\Delta s=1$ S-wave hyperon amplitudes within their
uncertainties.
In this way we determine that, with $\Lambda_\chi = 1$ GeV,
\begin{eqnarray}
{\cal A}_0^{(S)} (p \bar p K^0) + {\cal A}_1^{(S)} (p \bar p
K^0) & = &
1.5\pm 0.1 \nonumber \\
{\cal A}_0^{(S)} (p \bar n K^+) + {\cal A}_1^{(S)} (p \bar n
K^+) & = &
0.4\pm 0.1 \nonumber \\
{\cal A}_0^{(S)} (n \bar n K^0) + {\cal A}_1^{(S)} (n \bar n
K^0) & = &
1.9\pm 0.1
\end{eqnarray}
The SU(3) corrections tend to suppress the couplings compared to
their
tree level values,
a significant contribution of which comes from the use of the
one-loop
extracted weak
couplings in the tree level amplitudes instead of those extracted
at
tree level.
The corrections to the $ppK$ and $nnK$ couplings are, as one
expects for
SU(3) breaking,
at the $30\%$ level.
However, the corrections to the $pnK$ amplitude lead to an
effective coupling
approximately half the strength computed at tree level.
\subsection{P-Waves}
The tree level P-wave amplitudes come directly from
pole graphs involving one weak vertex from Eq.~\ref{weakl} and
one strong vertex from Eq.~\ref{strongl}, (see Fig.~\ref{tree})
\begin{eqnarray}\label{ptree}
{{\cal A}_0^{(P)} (p \bar p K^0) \over \Lambda_\chi} &=&
-{(D-F)(h_D-h_F)
\over m_N-m_\Sigma}
\nonumber \\
{{\cal A}_0^{(P)} (p \bar n K^+) \over \Lambda_\chi} &=& -{1 \over
6}{(D+3F)(h_D+3h_F)
\over m_N-m_\Lambda} + {1 \over
2}{(D-F)(h_D-h_F)
\over m_N-m_\Sigma}\nonumber \\
{{\cal A}_0^{(P)} (n \bar n K^0) \over \Lambda_\chi} &=& -{1 \over
6}{(D+3F)(h_D+3h_F)
\over m_N-m_\Lambda}-{1 \over 2}{(D-F)(h_D-h_F)
\over m_N-m_\Sigma}
\ \ \ \ .
\end{eqnarray}
\noindent The numerical values for these amplitudes
are found by using the parameters (extracted
from tree level comparison to experiment)
$h_D$ = --0.58, $h_F$ = 1.40, $D$ = 0.8, and $F$ = 0.5 \cite{ej,man}.
This yields tree level P-amplitudes of --2.4, 9.1, and 6.7
$(\times \Lambda_\chi / 1 \, {\rm GeV})$, respectively.
Loop diagrams as shown in Fig.~\ref{pwaves} give
\begin{eqnarray}\label{plooppp}
{{\cal A}_1^{(P)} (p \bar p K^0) \over \Lambda_\chi} &=&
{m_K^2 \over 16\pi^2 f_K^2} \ln\left({m_K^2 \over
\Lambda_\chi^2}\right) \times \nonumber\\
& & \left[
{ h_D \over m_N-m_\Sigma}
\left[ {10\over 3}D - {10\over 3}F + {92\over 9}D^3 - {140\over
9}D^2F
+{20\over 3}DF^2 - 4 F^3 \right] \right. \nonumber\\
& & \left.
+ { h_F \over m_N-m_\Sigma}
\left[ {10\over 3}F - {10\over 3}D - {56\over 9} D^3 +
{140\over 9}D^2F
- {32\over 3}DF^2 + 4 F^3 \right]
\right.\nonumber\\
& & \left.
+ h_\pi \left[ {5 \over 36} D + {7 \over 36}F - {41 \over 36}D^3
- {65 \over 36}D^2F + {313 \over 36}DF^2 -
{37 \over 12}F^3 - {5 \over 6}C^2(D-F)\right]
\right]
\nonumber\\
&-& {10 \over 81} C^2 {\cal H} {h_D-h_F\over m_N-m_\Sigma}
{\cal J}\big(\Delta m^{\Sigma^*}_N\big)
- {2 \over 9} {D-F \over m_N-m_\Sigma} C^2 h_C
{\cal J}\big(\Delta m^{\Sigma^*}_N\big) \nonumber \\
&-&{2 \over 9} C^2{h_D-h_F \over m_N - m_\Sigma}
\big[ (D + 5F) {\cal K}(m_K,\Delta m^{\Sigma^*}_N)
-(D-3F){\cal K}(m_\eta,\Delta m^{\Sigma^*}_N) \big] \nonumber \\
&+& {10 \over 81}h_\pi C^2 {\cal H} {\cal J}(\Delta m^{\Delta}_N)
\nonumber \\
&+& h_\pi C^2\Big[{4 \over 9}(D-F){\cal K}(m_K,\Delta m^\Delta_N)
+{2 \over 9}(F+D){\cal K}(m_K,\Delta m^{\Sigma^*}_N)
\nonumber \\
& & \phantom{ h_\pi C^2\Big[}
-{1\over 27}(D-3F)G_{\eta K\Sigma^*} - {1 \over 12}(D-F)
\tilde{G}_{\eta K \Sigma^*} \Big]
\nonumber \\
&+& {(D-F)(h_D-h_F) \over m_N-m_\Sigma} {14\over 3}C^2{\cal
J}(\Delta
m_\Sigma)
-{\cal A}_0^{(P)}(p \bar p K^0){\cal Z}_\Psi
\ \ \ ,
\end{eqnarray}
\noindent where ${\cal Z}_\Psi$ is the same wavefunction
renormalization
employed for the S-wave expressions.
\begin{eqnarray}\label{plooppn}
{{\cal A}_1^{(P)} (p \bar n K^+) \over \Lambda_\chi} &=&
{m_K^2 \over 16\pi^2 f_K^2} \ln ({m_K^2 \over \Lambda_\chi^2})
\Big({h_D \over m_N-m_\Lambda}
\Big[{5\over 9}D + {5\over 3}F + {2\over 27}D^3 + 2 D^2F
+ {34 \over 3}DF^2 + 6F^3\Big] \nonumber \\
&& \hskip 1.1cm + {h_F \over m_N-m_\Lambda}
\Big[{5\over 3}D + 5F + {20 \over 9}D^3 + 6D^2F
+ 16 DF^2 + 18 F^3\Big] \nonumber \\
&& \hskip 1.1cm + {h_D \over m_N-m_\Sigma}
\Big[ - {5\over 3}D + {5\over 3}F - {46 \over 9}D^3 + {70
\over 9}D^2F
- {10\over 3} DF^2 + 2 F^3\Big] \nonumber \\
&& \hskip 1.1cm + {h_F \over m_N-m_\Sigma}
\Big[{5\over 3}D - {5\over 3}F + {28 \over 9}D^3 - {70
\over 9}D^2F
+ {16 \over 3}DF^2 - 2 F^3\Big] \nonumber
\\
& & \hskip 1.1cm - h_\pi \Big[{11 \over 36}D + {11 \over 36}F -
{101 \over 108}D^3- {23\over 12}D^2F + {23\over 36}DF^2 -
{17\over 12}F^3 + C^2({1 \over 24} F-{13 \over 24} D)
\Big] \Big) \nonumber\\
& + & {5 \over 9}C^2{\cal H} {\cal J}(\Delta m_N^{\Sigma^*})
\Big[{1 \over 3}{h_D+3h_F \over m_N-m_\Lambda}+
{1 \over 9}{h_D-h_F\over m_N-m_\Sigma}\Big]
\nonumber \\
&+&C^2\Big[\Big(-{1 \over 3}{(h_D+3h_F)(D+F) \over
m_N-m_\Lambda}+
{1 \over 9} {(h_D-h_F)(D+5F) \over
m_N-m_\Sigma}\Big)
{\cal K}(m_K,\Delta m_N^{\Sigma^*}) \nonumber \\
&& \hskip 1.1 cm-{1 \over 9}{(h_D-h_F)(D-3F) \over
m_N-m_\Sigma}
{\cal K}(m_\eta,\Delta m_N^{\Sigma^*})\Big] \nonumber \\
&+& C^2 h_C {\cal J}(\Delta m_N^{\Sigma^*})
\Big[{1 \over 9}{D-F \over m_N-m_\Sigma} + {1 \over 3}
{D+3F \over m_N-m_\Lambda}\Big]
+ {10 \over 81} h_\pi C^2 {\cal H} {\cal J}(\Delta
m_N^{\Sigma^*})
\nonumber \\
&+&h_\pi C^2\Big[{4 \over 9}(D-F){\cal K}(m_K,\Delta m_N^\Delta)-
{5 \over 18}(F+D){\cal K}(m_K,\Delta m_N^{\Sigma^*})
\nonumber \\
& & \phantom{ h_\pi C^2\Big[}
-{1\over 54}(D-3F)G_{\eta K\Sigma^*}+
{1 \over 24}(D-F) \tilde{G}_{\eta K \Sigma^*} \Big]
\nonumber \\
&+& C^2 \Big[{1 \over 3}{(h_D+3h_F) \over m_N-m_\Lambda}(D+3F)
-{7 \over 3} {h_D-h_F \over m_N-m_\Sigma}(D-F) \Big]
{\cal J}(\Delta m_\Sigma) \nonumber \\
&-&{\cal A}_0^{(P)}(n \bar p K^+){\cal Z}_\Psi
\ \ \ \ ,
\end{eqnarray}
\begin{eqnarray}\label{ploopnn}
{{\cal A}_1^{(P)} (n \bar n K^0) \over \Lambda_\chi} &=&
{m_K^2 \over 16\pi^2 f_K^2} \ln ({m_K^2 \over \Lambda_\chi^2})
\Big({h_D \over m_N-m_\Lambda}
\Big[{5\over 9}D + {5\over 3}F + {2 \over 27}D^3
+ 2 D^2F
+ {34 \over 3}DF^2 + 6 F^3\Big] \nonumber
\\
&& \hskip 1.1cm + {h_F \over m_N-m_\Lambda}
\Big[ {5\over 3}D + 5 F + {20\over 9}D^3 +
6D^2F
+ 16 DF^2 + 18 F^3\Big] \nonumber \\
&& \hskip 1.1cm + {h_D \over m_N-m_\Sigma}
\Big[{5\over 3}D - {5\over 3}F + {46 \over 9}
D^3
- {70\over 9}D^2 F + {10 \over 3}DF^2 - 2
F^3\Big]
\nonumber \\
&& \hskip 1.1cm +{h_F \over m_N-m_\Sigma}
\Big[-{5\over 3}D + {5\over 3}F - {28 \over 9}D^3
+ {70 \over
9}D^2F
-{16 \over 3}DF^2 + 2 F^3\Big]
\nonumber \\
&& \hskip 1.1 cm - h_\pi \Big[{1 \over 6} D + {1 \over 9}F
+{11\over 54}D^3 - {1\over 9} D^2F - {145 \over 18} DF^2
+ {5\over 3}F^3 + C^2({7 \over 24}D-{19\over 24}F)\Big]\Big)
\nonumber\\
&+& {5 \over 9}C^2{\cal H} {\cal J}(\Delta m_N^{\Sigma^*})
\Big[{1 \over 3}{h_D+3h_F\over m_N-m_\Lambda}-
{1 \over 9}{h_D-h_F\over m_N-m_\Sigma}\Big]
\nonumber \\
&+&C^2\Big[\Big(-{1 \over 9}{(h_D-h_F)(D+5F) \over m_N-m_\Sigma}
-{1 \over 3}{(h_D+3h_F)(D+F) \over m_N -
m_\Lambda}\Big)
{\cal K}(m_K,\Delta m_N^{\Sigma^*}) \nonumber \\
&& \hskip 1.1cm +\Big({1 \over 9}{(h_D-h_F)(D-3F) \over
m_N-m_\Sigma}\Big)
{\cal K}(m_\eta,\Delta m_N^{\Sigma^*})\Big] \nonumber \\
&-& C^2 h_C {\cal J}(\Delta m_N^{\Sigma^*})
\Big[{1 \over 9}{D-F\over m_N-m_\Sigma}
-{1 \over 3}{D+3F\over m_N-m_\Lambda}\Big]
+ {20 \over 81}h_\pi C^2{\cal H} {\cal J}(\Delta m_N^{\Delta})
\nonumber \\
&+& h_\pi C^2 \Big[
{8 \over 9}(D-F){\cal K}(m_K,\Delta_N^\Delta)-
{1 \over 18}(F+D){\cal K}
(m_K,\Delta_N^{\Sigma^*})
\nonumber \\
& & \phantom{ h_\pi C^2 \Big[}
-{1\over 54}(D-3F) G_{\eta K \Sigma^*}-{1 \over 24}(D-F)
\tilde{G}_{\eta K \Sigma^*} \Big]
\nonumber \\
&+& C^2 \Big[{1 \over 3}{h_D+3h_F \over m_N-m_\Lambda}(D+3F)+
{7 \over 3}{h_D-h_F\over m_N-m_\Sigma}(D-F)\Big]{\cal J}(\Delta
m_\Sigma)
\nonumber \\
&-&{\cal A}_0^{(P)}(n \bar n K^0){\cal Z}_\Psi
\ \ \ \ .
\end{eqnarray}
\noindent The function ${\cal K}(m,\Delta)$ which appears in
Eq.~\ref{plooppp},
Eq.~\ref{plooppn} and Eq.~\ref{ploopnn}
from diagrams having both decuplet and octet intermediate states
is
\begin{eqnarray}
{\cal K}(m,\delta) = {1 \over 16\pi^2 f_K^2} & & \Big\{
(m^2 - {2\over 3}\delta^2) \ln({m^2 \over \Lambda_\chi^2})
\nonumber\\
& & +{2 \over 3}{1 \over \delta}\Big[(\delta^2-m^2)^{3/2}
\ln\left( {\delta-\sqrt{\delta^2-m^2+i\epsilon} \over
\delta+\sqrt{\delta^2-m^2+i\epsilon}}\right)
+ \pi m^3 \Big]\Big\}
\ \ \ ,
\end{eqnarray}
\noindent and the functions $G_{m_1,m_2,B}$ and
$\tilde{G}_{m_1,m_2,B}$ are given by
\begin{eqnarray}
G_{m_1,m_2,B} & = & {m_1^2 \over m_1^2-m_2^2}
{\cal K}(m_1,\Delta_N^B) + {m_2^2 \over m_2^2-m_1^2}
{\cal K}(m_2,\Delta_N^B)\ \ , \nonumber \\
\tilde{G}_{m_1,m_2,B} & = & {m_1^2 \over m_1^2-m_2^2}
{\cal J}(m_1,\Delta_N^B) + {m_2^2 \over m_2^2-m_1^2}
{\cal J}(m_2,\Delta_N^B)\ \ \ \ .
\end{eqnarray}
\noindent The mass differences that appear in Eq.~\ref{plooppp},
Eq.~\ref{plooppn}
and
Eq.~\ref{ploopnn}\ are defined by
\begin{eqnarray}
\Delta m^{\Sigma^*}_N & = & m_\Sigma^* - m_N \ \ ,\ \
\Delta m^\Delta_N = m_\Delta-m_N \nonumber\\
\Delta m_\Sigma & = & m_{\Xi^*} - m_\Sigma \ \sim \ \Delta
m^{\Sigma^*}_N
\ \ \ .
\end{eqnarray}
In the same way that the S-wave KNN amplitudes and associated
uncertainties
were determined, we use the expressions for S-wave hyperon decay
\cite{ej},
along with experimental measurements, to generate P-wave KNN
amplitudes
consistent
with S-wave hyperon decay rates.
The results, with $\Lambda_\chi=1$ GeV, are
\begin{eqnarray}
{\cal A}_0^{(P)} (p \bar p K^+) + {\cal A}_1^{(P)} (p \bar p
K^+) & = &
-1.7\pm 0.2 \nonumber \\
{\cal A}_0^{(P)} (p \bar n K^+) + {\cal A}_1^{(P)} (p \bar n
K^+) & = &
7\pm 1 \nonumber \\
{\cal A}_0^{(P)} (n \bar n K^+) + {\cal A}_1^{(P)} (n \bar n
K^+) & = &
6\pm 1 \nonumber \\
\end{eqnarray}
The P-wave amplitudes are seen to be reduced by $\sim 30\%$ from
their tree level values
by the SU(3) breaking
one-loop contributions. This is in contrast to the $\Delta s=1$
hyperon decay P-wave
amplitudes, where the corrections are at the $100\%$ level.
The NNK SU(3) breaking is the size one would naively guess and is
consistent with the idea
\cite{ej} that the large corrections to the P-wave amplitudes
for $\Delta s=1$
hyperon decays are the result of accidentally small tree level
amplitudes,
and not a breakdown of chiral perturbation theory.
Table I summarizes our findings. \footnote{We thank C. Bennhold for pointing
out a factor of 2 error in the original numerical values given for
the P-waves in the table. This is corrected in the table shown.}
\section{Discussion}
Weak NNK amplitudes that contribute to nonmesonic hypernuclear
decay
are not directly measurable but can be related to $\Delta s=1$
hyperon
decay by flavour SU(3).
We have computed the leading SU(3) breaking contributions to
these amplitudes
using heavy baryon chiral perturbation theory and find that such
corrections can suppress both the S-wave and P-wave amplitudes
by up to $50\%$.
\begin{table}
\begin{tabular}{ccccc}
& \multicolumn{2}{c}{\em S-waves} & \multicolumn{2}{c}{\em P-waves} \\
{\em vertex}
& ${\cal A}^{(S)}_0$ & ${\cal A}^{(S)}_0+{\cal A}^{(S)}_1 $
& ${\cal A}^{(P)}_0$ & ${\cal A}^{(P)}_0 + {\cal A}^{(P)}_1$
\\ \tableline
\rule{0cm}{0.5cm}
$p{\overline p}K^0$
& 2.0
&$ 1.5\pm 0.1 $
&$-2.4$
&$-1.7\pm 0.2 $
\\ \tableline
\rule{0cm}{0.5cm}
$p{\overline n}K^+$
&0.8
&$ 0.4\pm 0.1 $
&9.1
&$ 7\pm 1 $
\\ \tableline
\rule{0cm}{0.5cm}
$n{\overline n}K^0$
&2.8
&$ 1.9\pm 0.4 $
&6.7
&$ 6\pm 1 $
\\
\end{tabular}
\vskip 0.5cm
\caption{The S-wave and P-wave amplitudes at tree level and at
one-loop. We have set $\Lambda_\chi = 1$ GeV in both the S-wave and
P-wave amplitudes.}
\end{table}
In $\Delta s=1$ mesonic hyperon decay there are two tree level
graphs
contributing to P-wave amplitudes which tend to cancel against
each other
for the values of weak
couplings constants determined from S-wave hyperon decay
amplitudes.
Since there is only one graph contributing to the P-wave NNK
vertices, no
such
cancellations arise and the amplitudes are, in general, less
susceptible
to large SU(3) violation.
It would be interesting to compare the P-wave amplitudes
extracted from
hypernuclear decay with the amplitudes computed in this work.
It would help us to determine if the disagreement between the
observed
and predicted
$\Delta s=1$ mesonic hyperon decay P-wave amplitudes is an
accident of
nature or a hint that chiral perturbation theory is not
applicable to these
processes.
We stress that our computation is only an estimate of SU(3)
breaking effects
as there are unknown counterterms that also contribute.
The computation that we have performed is the leading effect in
the chiral
limit, $m_q\rightarrow 0$.
There is no reason to suspect that the counterterms cancel the
loop contributions since
the counterterms arise from UV physics whereas the nonanalytic
terms from the loop graphs are IR effects.
In order to determine the impact of our work on the understanding
of hypernuclear decay
the NNK amplitudes, including the SU(3) breaking
corrections,
must be incorporated into a realistic hypernucleus in the same
way that
previous estimates of interaction strengths have been included,
e.g. \cite{RMBJ92}.
It seems likely that the results found in this work will have
significant impact on theoretical predictions for the ratio
of neutron-induced to proton-induced decay widths of
$\Lambda$-hypernuclei, eg. $^{12}_\Lambda C$ \cite{benntalk95},
and possibly other flavours of hypernuclei.
\section{Acknowledgements}
We would like to thank the Institute for Nuclear Theory at the
University of
Washington for their kind hospitality during some of this work.
RPS would also like to thank the Institute for Theoretical
Physics at the
University of California, Santa Barbara, and Carnegie-Mellon
University, where some of this work was completed.
MJS would like to thank Duke University where some of this work
was carried out.
MJS would also like to thank L. Kisslinger and C. Bennhold for
useful discussions.
Our work is supported in part by the US Dept. of Energy under
grant number
DE-FG02-91-ER40682, and grant number DE-FG05-90ER40592.
|
\subsection*{Abstract}
A numerical study of static, spherically symmetric
monopole solutions of a spontaneously broken SU(2) gauge theory
coupled to a dilaton field is presented.
Regular solutions seem to exist only up a maximal value of
the dilaton coupling.
In addition to the generalization of the 't~Hooft-Polyakov
monopole a discrete family of regular solutions is found,
corresponding to radial excitations absent in the theory
without dilaton.
\endgroup
\newpage
\noindent
The aim of this paper is to present a detailed numerical
study of classical solutions of an
SU(2) Yang-Mills-Higgs (YMH) theory coupled to a
(massless) dilaton (DYMH),
when the Higgs field is in the adjoint representation.
Dilaton fields appear naturally in low energy effective
field theories derived from superstring models \cite{Wi, BFQ, GSW}.
As previous studies have already shown \cite{LavI, BizI} the
inclusion of a dilaton in a pure Yang-Mills (YM) theory
has drastic consequences already at the classical level.
In particular the dilaton Yang-Mills (DYM)
theory possesses finite energy `particle-like'
solutions which are absent in the pure YM case.
The same phenomenon happens in the
Einstein-Yang-Mills system where non-singular finite
energy solutions have been discovered some time ago \cite{BM}.
The above YMH model without a dilaton field
is known to possess nonsingular, finite energy
solutions, describing magnetic monopoles \cite{THP}.
In the present work we study static, spherically symmetric
solutions of the DYMH theory.
Our results show a striking similarity to those obtained
in the corresponding YMH theory coupled to gravity (EYMH)
\cite{BFM}.
In the DYMH theory
there is a finite energy abelian solution for all values of
the dilaton coupling {$\alpha$}.
Based on numerical investigations there is a strong
indication for the existence of a finite energy
{\sl nonabelian} monopole up to only
a maximal value of $\alpha$, $\alpha_{\rm m}$.
The nonabelian monopole merges with the abelian
one at a critical value of $\alpha$, $\alpha_{\rm c}$.
In limit when the dilaton decouples the nonabelian
solution joins smoothly
the 't Hooft-Polyakov monopole. If the Higgs self
coupling, $\beta$, is
in the interval [0,0.6] $\alpha_{\rm m}(\beta)$ and
$\alpha_{\rm c}(\beta)$
are different and with close analogy to the EYMH theory there are two
different solutions if $\alpha_{\rm c}<\alpha<\alpha_{\rm m}$.
In addition to the `fundamental' monopole there is
good numerical evidence that
a countable family
of globally regular solutions exits
for $0<\alpha<\sqrt3/2$ independently of
$\beta$.
As their energy is
higher they can be interpreted as excitations of the
fundamental monopole.
In the limit when the dilaton filed decouples
their energy diverges.
There is another limit when the DYMH theory reduces to the
DYM model and
then these excitations tend to the solutions discovered
in refs.~\cite{LavI,BizI}.
The action of our model is given by
\begeq{sdymh}
S=\int d^4x\left\{{1\over2}\partial_\mu\varphi\,\partial^\mu\varphi-
e^{2\kappa \,\varphi}{1\over4g^2}F_{\mu\nu}^aF^{a\mu\nu}+
{1\over 2}(D_\mu\Phi)^a (D^\mu\Phi)^a-e^{-2\kappa \,\varphi}\,V(\Phi)
\right\}\,,
\end{equation}
where $a=1$,$2$,$3$, and the Higgs potential is
\begeq{higgspot}
V(\Phi)={\lambda\over8}(\Phi^a\Phi^a-v^2)^2\,.
\end{equation}
The couplings of the scalar field $\varphi$ has been chosen so that the
action (\ref{sdymh}) possesses the following dilatational symmetry
\begeq{dymhdil}
x^\mu\to e^{\kappa c}x^\mu\;,\quad A_\mu\to e^{-\kappa c}A_\mu\;,
\quad \varphi\to\varphi+c\;,\quad S\to e^{2\kappa c}S\,,
\end{equation}
which can be used to fix the value of $\varphi$ arbitrarily at any point.
With the
\begeq{sscale}\eqalign{
\Phi\to v\Phi,\quad x\to{1\over gv}x,&\quad\varphi\to v\varphi,
\quad S\to{v\over g}S,\cr
\lambda\to{g^2}\beta^2,&\quad\kappa \to{1\over v}\alpha}
\end{equation}
rescaling of variables the only remaining dimensionless parameters are
\begeq{alphabeta}
\alpha={\kappa M_W\over g}\;,\quad \beta={M_H\over M_W}\;,
\end{equation}
where $M_W=vg$ and $M_H=v\sqrt{\lambda}$ are the
characteristic masses of the theory.
Without further restrictions one can assume $\alpha\geq0$.
Let us mention here that the action of the DYMH theory (\ref{sdymh})
can be obtained from the Einstein-Yang-Mills-Higgs
action by restricting the
metric to the special conformstatic form
$ds^2=e^{2\phi}dt^2-e^{-2\phi}dx^idx^i$.
We use the minimal static spherically symmetric ansatz with zero
`electric' field
\begin{subeqnarray}\label{minanz}
A^a_0\equiv 0,\quad
A^a_i&=&\epsilon_{aij}{x^j\over r^2}(W(r)-1)\;,\\
\Phi^a&=&H(r){x^a\over r}\;,\\
\varphi&=&\varphi(r)\;.
\end{subeqnarray}
After the rescaling (\ref{sscale}) the action (\ref{sdymh}) reduces
to
\begeq{sdymhanz}
E=\int dr\left\{{1\over2}r^2{\varphi'}^2+e^{2\alpha\varphi}
\Bigl[
{W'}^2+{(W^2-1)^2\over2r^2}\Bigr]+W^2H^2+
{r^2{H'}^2\over2}+
{\beta^2\over8}r^2e^{-2\alpha\varphi}(H^2-1)^2\right\}
\end{equation}
using the ansatz (\ref{minanz}).
Varying (\ref{sdymhanz}) we obtain the equations of motion:
\begin{subeqnarray}\label{dymhanzeq}
\Bigl(r^2\varphi'\Bigr)'&=&2\alpha e^{2\alpha\varphi}\Bigl({W'}^2+
{(W^2-1)^2\over2r^2}\Bigr)-{\alpha\beta^2\over4}r^2e^{-2\alpha\varphi}
(H^2-1)^2\;,\\
r^2W''&=&W(W^2-1+r^2e^{-2\alpha\varphi}H^2)-2\alpha r^2\varphi'W'\;,\\
(r^2H')'&=&H\Bigl(2W^2+{\beta^2\over2}r^2e^{-2\alpha\varphi}(H^2-1)\Bigr)
\;.
\end{subeqnarray}
Solutions regular at the origin must satisfy the following boundary
conditions (b.c.):
\begeq{dymhsor}\eqalign{
H=ar+O(r^2),\quad W&=1-br^2+O(r^3),\quad \cr
\varphi=\varphi_0+\alpha\Bigl(2b^2e^{2\alpha\varphi_0}&-
{\beta^2\over24}e^{-2\alpha\varphi_0}\Bigr)r^2+O(r^3)\;,}
\end{equation}
i.e.~there is a three parameter ($a$, $b$, $\varphi_0$)
family of regular solutions at $r=0$. The local existence can be proved
following the procedure discussed in \cite{BFMII}
For $r\to\infty$ the corresponding `regular' b.c. are
\begin{subeqnarray}\label{dymhinfreg}
\varphi&=&\varphi_\infty-{d\over r}+O({1\over r^2})\;,\\
W&=&B e^{-\mu\rho}\left(1+O({1\over r})\right)\;,\\
H&=&1-{C\over r}e^{-\mu\beta\rho}\left(1+O({1\over r})
\right)\quad{\rm for}\;\;\beta<2\\
H&=&1-{2B^2\over {\mu^2(\beta^2-4)r^2}}e^{-2\mu\rho}
\left(1+O({1\over r})\right)\quad{\rm for}\;\;\beta>2\;,
\end{subeqnarray}
parametrized by ($\varphi_\infty$, $d$, $B$, $C$), where
$\mu=e^{-\alpha\varphi_\infty}$, $\rho=r+\alpha d \ln r$.
For $\beta>2$ we cannot fully parametrize the stable
manifold at $r=\infty$.
Exploiting the virial theorem and using Eqs.~(\ref{dymhanzeq}),
(\ref{dymhsor}) and (\ref{dymhinfreg}) one obtains for the energy
(\ref{sdymhanz})
\begeq{dymhvire}
E={1\over\alpha}\int\limits_0^\infty dr\,(r^2\varphi')'=
{1\over\alpha}\left.(r^2\varphi')\right|_0^\infty.
\end{equation}
We discuss first an exact two parameter family of solutions of
Eqs.~(\ref{dymhanzeq}) which is going to play
an important role in our further analysis,
since it descibes the asymptotic behaviour of nonabelian solutions.
Consider the singular abelian monopole, $W\equiv0$, $H\equiv1$ then
the general solution of (\ref{dymhanzeq}) takes the form
\begeq{dymhgenrn}
\varphi_{\rm a}=-{1\over\alpha}\ln\abs{{1\over A}\sinh\Bigl(A({1\over M}+
{\alpha\over r})\Bigr)}
\end{equation}
with $A^2,M\in{\bf R}$.
For $M>0$ and $A$ real it is regular for all $r>0$.
Its energy, $E$ Eq.~(\ref{sdymhanz}), is infinite unless $A=0$. For $A=0$
the total energy of the solution (\ref{dymhgenrn}) is finite, $E=M/\alpha$.
When $M<0$ and $A$ is real the solution is singular at $r=-\alpha M$ and
the energy diverges.
If $A$ is imaginary for any value of $M$ the solution becomes singular at
some $r>0$.
Using the dilatational symmetry (\ref{dymhdil}) from now on we set
$\varphi_\infty=0$. Note that then the energy scales as
$E\to \exp(-\alpha\varphi_\infty)E$. The finite energy abelian solution
for $\varphi_\infty=0$ reads
$$\varphi=-{1\over\alpha}\ln(1+{\alpha\over r})\,,$$
and its energy is simply $E=1/\alpha$.
It satisfies the following first order differential Eq.
\begeq{dymhbogeq}
r^2\varphi'=\pm e^{\alpha\varphi}\,,
\end{equation}
which is derivable from a Bogomolny type bound \cite{BizI} in the
background $W\equiv0$, $H\equiv1$.
The asymptotic behaviour of finite energy nonabelian solutions is described
by (\ref{dymhgenrn}). With the help of Eq.~(\ref{dymhvire})
we arrive at the following useful formula for the total energy
of globally regular nonabelian solutions:
\begeq{dymhaszenergscal}
E={1\over\alpha}\cosh(A/M)\;,
\end{equation}
where $A$, $M$ are parameters of $\varphi_{\rm a}$ describing their
asymptotic behaviour.
For globally regular solutions of (\ref{dymhanzeq}) satisfying
the b.c.~(\ref{dymhsor}), (\ref{dymhinfreg}) the energy (\ref{sdymhanz})
is a monotonically increasing function of the Higgs coupling $\beta$ and
monotonically decreases with the dilaton coupling $\alpha$.
(We remark here that the condition $\varphi_\infty=0$ is
essential to prove these facts.)
We have numerically integrated a suitably desingularized version of
Eqs.~(\ref{dymhanzeq}) from $r=0$ using the b.c.~(\ref{dymhsor}).
We set $\varphi_0$ to zero in order to have to vary only two parameters.
By adjusting the parameters $a$, $b$ we suppressed the divergent modes
in the asymptotic region according to (\ref{dymhaszmod}). The resulting
solutions can then be transformed to satisfy $\varphi_\infty=0$ by the
dilatational symmetry (\ref{dymhdil}).
We found that there exist a fundamental nonabelian monopole solution
for all $\beta$ below a maximal value of $\alpha$, $\alpha_{\rm m}$
which smoothly joins
the 't~Hooft-Polyakov monopole in the $\alpha\to 0$ limit.
Some solution curves for $\beta=0$ are shown in Fig.~\ref{0b0c}.
\begin{figure}
\hbox to\hsize{\hss
\figbox{0.5\hsize}{70 72 818 556}{0bwh}\hss
\figbox{0.5\hsize}{70 72 818 556}{0bdil}\hss }
\begin{picture}(0,0)(0,0)
\put(205,0){\makebox(0,0){\footnotesize$\ln(1+r)$}}
\put(435,0){\makebox(0,0){\footnotesize$\ln(1+r)$}}
\put(25,170){\makebox(1,1){\small\footnotesize$W$, \footnotesize$H$}}
\put(250,170){\makebox(1,1){\small$\varphi$}}
\end{picture}
\caption[0b0c]{\label{0b0c}
The $\alpha$ dependence of the fundamental monopole solution for
$\beta=0$}
\end{figure}
For the parameters see Table~{\ref{T1}}.
\begin{table}[htb]
\begin{minipage}[t]{10.3cm}
\caption{}
\vspace{.2truecm}
\label{T1}
{\small
\begin{tabular}{|l|c|c|l|c|}
\hline
\multicolumn{5}{|c|}{Parameters for the fundamental monopole
solutions in Fig.~\ref{0b0c}}\\ \hline
\multicolumn{1}{|c|}{$\alpha$}&
\multicolumn{1}{|c|}{$a$}&
\multicolumn{1}{|c|}{$b$}&
\multicolumn{1}{|c|}{$\phi_\infty$}&
\multicolumn{1}{|c|}{$\alpha\cdot E$}\\ \hline\hline
0.0 & 0.33333333 & 0.16666666 & 0 & 0{\phantom .}~~~~~~~~~ \\
0.1 & 0.33324422 & 0.16674419 & 0.050094 & 0.099853 \\
0.2 & 0.33297626 & 0.16697893 & 0.100758 & 0.198827 \\
0.5 & 0.33107418 & 0.16872021 & 0.262673 & 0.481732 \\
0.7 & 0.32884174 & 0.17095758 & 0.387902 & 0.650049 \\
0.8 & 0.32741363 & 0.17252229 & 0.460048 & 0.725582 \\
1.0 & 0.32388447 & 0.17699856 & 0.638684 & 0.855180 \\
1.1 & 0.32176986 & 0.18029088 & 0.758343 & 0.907479 \\
1.2 & 0.31944319 & 0.18485325 & 0.919110 & 0.950109 \\
1.3 & 0.31705890 & 0.19195967 & 1.172960 & 0.981707 \\
1.4 & 0.31657705 & 0.20987214 & 1.950448 & 0.999523 \\
1.4088 & 0.31898696 & 0.21924215 & 2.582925 & 1.000092 \\
1.4 & 0.32380520 & 0.22987901 & 5.289654 & 1.000000 \\
1.39939 & 0.32402298 & 0.23025275 & 8.461582 & 1.000000 \\
\hline
\end{tabular}
}
\end{minipage}
\hskip .5cm
\begin{minipage}[t]{5cm}
\caption{}
\vspace{.2truecm}
\label{T4}
{\small
\begin{tabular}{|l|c|r@{.}l|}
\hline
\multicolumn{1}{|c|}{$\beta ^2$}&
\multicolumn{1}{|c|}{$\alpha_{\rm c}$}&
\multicolumn{2}{|c|}{$\alpha_{\rm m}$} \\ \hline\hline
0 & 1.39938 & 1&4088 \\
0.02 & 1.37874 & 1&3803 \\
0.03 & 1.36950 & 1&3702 \\
0.04 & 1.36088 & 1&3612 \\
0.043 & 1.35839 & 1&3586 \\
0.05 & 1.35279 & 1&3529 \\
0.06 & $\alpha_{\rm m}$ & 1&3452 \\
0.07 & $\alpha_{\rm m}$ & 1&338 \\
0.08 & $\alpha_{\rm m}$ & 1&33 \\
0.1 & $\alpha_{\rm m}$ & 1&32 \\
0.15 & $\alpha_{\rm m}$ & 1&29 \\
0.2 & $\alpha_{\rm m}$ & 1&27 \\
1 & $\alpha_{\rm m}$ & 1&11 \\
4 & $\alpha_{\rm m}$ & 0&98\\
$\infty$ & $\alpha_{\rm m}$ & 0&89\\
\hline
\end{tabular}
}
\end{minipage}
\end{table}
There is a critical $\alpha$ value, $\alpha_{\rm c}$,
where the dilaton field
becomes logarithmically divergent while $W$ and $H$ tend to
some nontrivial
functions as can be seen on Fig.~\ref{0b0c}. After shifting
$\varphi_\infty$
to 0 the fundamental monopole tends to the finite energy abelian one as
$\alpha\to\alpha_{\rm c}$.
For $\beta^2\leq0.06$ $\alpha_{\rm c}$
is seen to differ from $\alpha_{\rm m}$
(see Table \ref{T4}).
It means that for a given $\alpha$ in the interval
[$\alpha_c,\alpha_{\rm m}$] there are two different nonabelian solutions
with different energies. The function
$\alpha_{\rm m}(\beta)$ decreases with increasing $\beta$ from
$\alpha_{\rm m}(0)\approx1.4088$ to $\alpha_{\rm m}(\infty)\approx0.89$.
There also seems to exist a countable family of globally regular monopole
solutions indexed by the number zeros of $W(r)$ for all $\beta$
and $0<\alpha<\sqrt3/2$. They can be interpreted as radial excitations of
the fundamental monopole. In the $\alpha\to 0$ limit after a suitable
rescaling they can be identified with the previously discovered DYM
solutions \cite{LavI},\cite{BizI}.
We illustrate some of these excited
solutions with one and two zeros for $\alpha$ various values
on Figs.~\ref{0b1c} and \ref{0b2c}, while the corresponding
parameters are listed in Table~\ref{T2} and Table~\ref{T3}.
\begin{table}[htb]
\centering
\caption{}
\vspace{.2truecm}
\label{T2}
\begin{tabular}{|l|l|l|r@{.}l|c|}
\hline
\multicolumn{6}{|c|}{Parameters for the first excited monopole
solution in Fig.~\ref{0b1c}}\\ \hline
\multicolumn{1}{|c|}{$\alpha$}&
\multicolumn{1}{|c|}{$a$}&
\multicolumn{1}{|c|}{$\alpha^2b$}&
\multicolumn{2}{|c|}{$\alpha\varphi_\infty$}&
\multicolumn{1}{|c|}{$\alpha\cdot E$} \\ \hline\hline
{\small DYM } & ~~~~~~---& 0.26083015 & 1&711412 & 0.8038078 \\
0.05 & 0.66718265 & 0.26286356 & 1&731264 & 0.8507008 \\
0.1 & 0.77749002 & 0.27004539 & 1&803886 & 0.8904364 \\
0.2 & 0.99147112 & 0.30099818 & 2&170909 & 0.9448188 \\
0.3 & 1.00351436 & 0.33435729 & 2&730953 & 0.9720677 \\
0.5 & 0.77902186 & 0.37296818 & 3&994232 & 0.9932192 \\
0.7 & 0.59081576 & 0.39281185 & 5&919563 & 0.9992733 \\
0.8 & 0.51698205 & 0.39916099 & 8&207917 & 0.9999547 \\
0.866& 0.47164657 & 0.40215328 & 26&31413 & 1.0000000 \\
\hline
\end{tabular}
\end{table}
\begin{table}[htb]
\centering
\caption{}
\vspace{.2truecm}
\label{T3}
\begin{tabular}{|l|l|l|l|c|}
\hline
\multicolumn{5}{|c|}{Parameters for the second excited monopole
solution in Fig.~\ref{0b2c}}\\ \hline
\multicolumn{1}{|c|}{$\alpha$}&
\multicolumn{1}{|c|}{$a$}&
\multicolumn{1}{|c|}{$\alpha^2b$}&
\multicolumn{1}{|c|}{$\alpha\varphi_\infty$}&
\multicolumn{1}{|c|}{$\alpha\cdot E$} \\ \hline\hline
{\small DYM} & ~~~~~~~--- & 0.35351801 & 3.373903 & 0.96559852 \\
0.05 & 6.0251373 & 0.36759603 & 4.169777 & 0.99251583 \\
0.1 & 3.9102367 & 0.37450939 & 4.996143 & 0.99673418 \\
0.2 & 2.0946637 & 0.37885368 & 5.999793 & 0.99882187 \\
0.3 & 1.4148625 & 0.38161925 & 6.815222 & 0.99949621 \\
0.5 & 0.85071037 & 0.38762800 & 8.666195 & 0.99993064 \\
0.7 & 0.60036405 & 0.39522735 & 12.25580 & 0.99999862 \\
0.8 & 0.51818563 & 0.39944391 & 17.77525 & 1.00000000 \\
0.84 & 0.48948129 & 0.40111050 & 25.77133 & 1.00000000 \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\hbox to\hsize{\hss
\figbox{0.5\hsize}{70 72 818 556}{0b1cwh}\hss
\figbox{0.5\hsize}{70 72 818 556}{0b1cdil}\hss }
\begin{picture}(0,0)(0,0)
\put(205,0){\makebox(0,0){\footnotesize$\ln(1+r/\alpha)$}}
\put(435,0){\makebox(0,0){\footnotesize$\ln(1+r/\alpha)$}}
\put(25,170){\makebox(0,0){\small\footnotesize$W$, \footnotesize$H$}}
\put(250,170){\makebox(0,0){\small$\alpha\varphi$}}
\end{picture}
\caption[0b1c]{\label{0b1c}
The $\alpha$ dependence of the first excited monopole solution for
$\beta=0$}
\end{figure}
\begin{figure}
\hbox to\hsize{\hss
\figbox{0.5\hsize}{70 72 818 556}{0b2cwh}\hss
\figbox{0.5\hsize}{70 72 818 556}{0b2cdil}\hss }
\begin{picture}(0,0)(0,0)
\put(205,0){\makebox(0,0){\footnotesize$\ln(1+r/\alpha)$}}
\put(435,0){\makebox(0,0){\footnotesize$\ln(1+r/\alpha)$}}
\put(25,170){\makebox(0,0){\small\footnotesize$W$, \footnotesize$H$}}
\put(250,170){\makebox(0,0){\small$\alpha\varphi$}}
\end{picture}
\caption[0b2c]{\label{0b2c}
The $\alpha$ dependence of the second excited monopole solution for
$\beta=0$}
\end{figure}
When $\alpha\to \sqrt3/2$ for the excited solutions the dilaton
diverges logarithmically again while the zeros of $W$ go rapidly to
infinity, except for the innermost one.
For $\sqrt3/2<\alpha<1$ there is a surviving solution with
a single zero of $W$ and with
divergent $\varphi$ existing up to $\alpha=1$, where $W$
develops an extra divergent mode, so for $\alpha>1$ this
solution is not expected to exist.
There also exists
another type of limiting solution when the number of zeros of $W$ goes to
infinity for all $0<\alpha<\sqrt3/2$ and the dilaton field diverges
logarithmically.
If one shifts, however $\varphi_\infty$ to 0 then all of the excited
monopoles merge with the finite energy abelian solution for
$\alpha\to\sqrt3/2$.
The fundamental and the excited monopoles up to six zeros are plotted on
Fig.~\ref{0ba2}. In order to better display the higher zeros of $W$ we
plotted $\sqrt r W$. Notice that the newer zeros appear nearer and
nearer to the origin.
\begin{figure}
\hbox to\hsize{\hss
\figbox{0.5\hsize}{70 72 818 556}{0ba2wh}\hss
\figbox{0.5\hsize}{70 72 818 556}{0ba2dil}\hss }
\begin{picture}(0,0)(0,0)
\put(200,0){\makebox(0,0){\small$\ln(r/\alpha)\!-\!\alpha\varphi_\infty$}}
\put(425,0){\makebox(0,0){\small$\ln(r/\alpha)\!-\!\alpha\varphi_\infty$}}
\put(25,170){\makebox(2,-1){\small$\sqrt r$\footnotesize$W$, \footnotesize$H$}}
\put(253,170){\makebox(2,-1){\small$\varphi\!-\!\varphi_\infty$}}
\end{picture}
\caption[0ba2]{\label{0ba2}
The fundamental and the first six excited monopole solutions for
$\alpha=0.2$ and $\beta=0$}
\end{figure}
The structure of the solutions can be understood from the linearization
of the field equations (\ref{dymhanzeq})
around the abelian monopole (\ref{dymhgenrn}). Using the
\begeq{dymhrnlinanz}
W=w(r)\;,\quad H=1+{h(r)\over r}\;,\quad
\varphi=\varphi_{\rm a}(r)+\psi\;,
\end{equation}
variables, from the linearized field equations one finds for
${1/M}\ll{\alpha /r}\ll{1/\abs A}$ (where
$\varphi_{\rm a}\simeq\ln(r/\alpha)/\alpha$) that the solutions are
well approximated by:
$\psi\sim e^{\lambda_\psi\tau},\; h\sim e^{\lambda_h\tau},\;
w\sim e^{\lambda_w\tau}$ where$\tau=\ln r$ and the `frequencies' are
\begeq{dymhimzmod}
\lambda_\psi=(1;-2)\;,\quad \lambda_h={1\over2}(1\pm
\sqrt{1+4\alpha^2\beta^2})\;,\quad
\lambda_w=-{1\over2}\pm\sqrt{\alpha^2-3/4}\,.
\end{equation}
So we see that in this `middle' region $W$ oscillates with an amplitude
decaying like $1/\sqrt r$.
If $M\to\infty$ (i.e. $\alpha E\to1$) this region streches out
to infinity, while
$W$ has more and more zeros when $\alpha<\sqrt3/2$.
This behaviour is a similar to the one found in the DYM case
in the interval
where $\varphi$ grows logarithmically.
If $\alpha>1$ $W$
we do not expect
the corresponding solution to exist.
In the asymptotic region defined by $r\gg\abs{\alpha A\coth(A/M)}$,
$r\gg\mu^{-1}$
(where $\mu=\abs{\sinh(A/M)/A}$ and
$\varphi_{\rm a}\simeq -{1\over\alpha}\ln\mu$)
the linear corrections $\beta\not=0$ are characterized
by $\psi\sim e^{\lambda_\psi\tau},\;w\sim e^{\lambda_w r},\;
h\sim e^{\lambda_h r}$, with
\begeq{dymhaszmod}
\lambda_\psi=(0;-1),\quad\lambda_h=\pm\mu\abs\beta,\quad
\lambda_w=\pm\mu,
\end{equation}
while $h\sim e^{\lambda_h\tau}$ and $\lambda_h=(0;+1)$ for $\beta=0$.
The energy of the solutions goes rapidly to $1$ if the number of
oscillations of $W$ increases (see Table~\ref{T5}).
\begin{table}[htb]
\centering
\caption{}
\vspace{.2truecm}
\label{T5}
\begin{tabular}{|l|l|l|r@{.}l|c|}
\hline
\multicolumn{6}{|c|}{Parameters of the solutions for $\alpha=0.2$ and
$\beta=0$ in Fig.~\ref{0ba2}}\\
\hline
\multicolumn{1}{|c|}{$n$}&
\multicolumn{1}{|c|}{$a$}&
\multicolumn{1}{|c|}{$b$}&
\multicolumn{2}{|c|}{$\varphi_\infty$}&
\multicolumn{1}{|c|}{$\alpha\cdot E$} \\ \hline\hline
0 & 0.33297626 & 0.16697893 & 0&1007578 & 0.1988268 \\
1 & 0.99147112 & 7.52495447 & 10&854544 & 0.9448188 \\
2 & 2.09466373 & 9.47134195 & 29&998964 & 0.9988219 \\
3 & 2.14396020 & 9.51995711 & 48&727339 & 0.9999720 \\
4 & 2.14545164 & 9.52111922 & 67&372640 & 0.9999993 \\
5 & 2.14549487 & 9.52114737 & 86&014682 & 1.0000000 \\
6 & 2.14549609 & 9.52114805 & 104&6566 & 1.0000000 \\
\hline
\end{tabular}
\end{table}
We have determined the energy of the solution by fitting the
parameters
$A$, $M$ in Eq.~(\ref{dymhgenrn}) in the asymptotic region
using
formula (\ref{dymhaszenergscal}). The energy determined
this way contains only exponentially small corrections.
We have also plotted the $\alpha$ dependence
of the energy, $E(\alpha)$, (rescaled to $\varphi_\infty=0$)
on Fig.~\ref{0bea0}
for $\beta=0$.
The similarity of Fig.~\ref{0bea0} to Fig~3 in Ref.~\cite{BFM}
where the masses of {\sl gravitating monopoles} are plotted as a
function of the `gravitational coupling strength',
$M_{\scriptstyle\rm W}/M_{\scriptstyle\rm Planck}$,
is indeed striking.
On Fig.~\ref{b02eaz} $E(\alpha)$ for $\beta^2=0.02$
and 0.043 is shown.
\begin{figure}
\hbox to\hsize{\hss
\figbox{0.5\hsize}{70 72 818 556}{0be-a012}\hss
\figbox{0.5\hsize}{70 72 818 556}{0be-a0z}\hss }
\begin{picture}(0,0)(0,0)
\put(205,0){\makebox(0,0){\small$\alpha$}}
\put(440,0){\makebox(0,0){\small$\alpha$}}
\put(25,165){\makebox(0,0){\small$\alpha$\footnotesize$E$}}
\put(250,165){\makebox(0,0){\small$\alpha$\footnotesize$E$}}
\end{picture}
\caption[0bea0]{\label{0bea0}
$\alpha$ dependence of the energy of the fundamental, the first, and
the second excited monopole solutions, and its
detailed view for the fundamental
monopole near $\alpha_{\rm m}$ for $\beta=0$}
\end{figure}
\begin{figure}
\hbox to\hsize{\hss
\figbox{0.5\hsize}{70 72 818 556}{b02e-az}\hss
\figbox{0.5\hsize}{70 72 818 556}{b043eaz}\hss }
\begin{picture}(0,0)(0,0)
\put(205,0){\makebox(0,0){\small$\alpha$}}
\put(440,0){\makebox(0,0){\small$\alpha$}}
\put(25,165){\makebox(0,0){\small$\alpha$\footnotesize$E$}}
\put(250,165){\makebox(0,0){\small$\alpha$\footnotesize$E$}}
\end{picture}
\caption[b02eaz]{\label{b02eaz}
Detailed view of the $\alpha$ dependence of the energy of the fundamental
monopole solution near $\alpha_{\rm m}$ for $\beta^2=0.02$ and for
$\beta^2=0.043$}
\end{figure}
For not too large $\beta$ values ($\beta\leq3$)
$\alpha E$ of the fundamental monopole becomes
larger than 1 unlike for the excited ones.
For $\beta=\infty$ this maximum is $1$.
We make finally some remarks on the stability of the solutions. The
't Hooft-Polyakov monopoles are stable since they are solutions with minimal
energy \cite{TFS}. It is natural to expect the fundamental monopoles to
remain stable in the DYMH case for $\alpha>0$. For sufficiently
small $\beta$, however,
where the mass of the fundamental monopole is larger than that of the abelian
one, the nonabelian solution is expected to become unstable against large
perturbations. If $\alpha_{\rm c}\not=\alpha_{\rm m}$ there is a bifurcation
point where the linear stability of the solutions can change. In the EYMH
case this change of stability has been shown in \cite{H}.
The excited monopoles are expected to be unstable for all $\alpha$ since
their energy is significantly larger than that of the fundamental ones.
This heuristic argument is strengthened by the fact that
in the $\alpha\to0$ limit their counterparts are known to be
unstable \cite{BizI,LavI}.
\newpage
|
\section{Introduction}
The possibility that space-time has more than four dimensions has
received much attention regarding its cosmological aspects
[1--8]. Investigations have focused on attempts to explain why the
universe presently appears to have only four space-time dimensions if it
is, in fact, a dynamically evolving $(4+k)$-dimensional manifold ($k$
being the number of extra dimensions). It has been shown that solutions
to the $(4+k)$-dimensional Einstein equations exist, for which
four-dimensional space-time expands while the extra dimensions contract
or remain constant [4--8]. It has been pointed out that the extra
dimensions can produce large amounts of entropy during the contraction
process \cite{AG}, thus providing an alternative resolution to the
horizon and flatness problems \cite{Guth}, as compared to the usual
inflationary scenario. It has been also suggested that experimental
detection of the time variation of the fundamental constants could
provide strong evidence for the existence of extra dimensions \cite{Mar}.
In the present paper we study the 5D cosmological model of
Kaluza-Klein type, with the fifth coordinate being a generalization of
the universal parametric ``historical'' time $\tau $ discussed by, for
example, Stueckelberg \cite{Stu} and Horwitz and Piron \cite{HP}. It has
been shown that gauge invariance of the Stueckelberg-Schr\"{o}dinger
equation requires the addition of a fifth gauge field \cite{SHA}; this
result also follows from Feynman's approach to the foundations of gauge
theories in a manifestly covariant framework \cite{LHS}. The equations of
motion for such a gauge field are of second order in the five-dimensional
manifold $(x^\mu ,\tau ),$ with metric (4,1) or (3,2); i.e., on the level
of the gauge fields, the parametric ``historical'' time has entered a
five-dimensional manifold, much in the way that the Newtonian time $t$
enters the four-dimensional Minkowski manifold as a consequence of the
requirements of full gauge invariance of the Schr\"{o}dinger equation.
The canonical quantization of this $U(1)$ gauge theory has been carried
out by Shnerb and Horwitz \cite{SH}, where it is shown that the standard
Maxwell theory is recovered in a ``correlation'' limit.
The present paper is concerned with a 5D theory of gravitation with
$\tau $ as a fifth coordinate, which originated in an earlier work
\cite{therm}, in which we considered the thermodynamics of a relativistic
$N$-body system, taking account of the mass distribution in such a system
\cite{BH1,ind}. In \cite{therm} we incorporated two-body interactions, by
means of the direct action potential $V(q),$ where $q$ is an invariant
distance in the Minkowski space, taking the support of the two-body
correlations to be in a $O(2,1)$ invariant subregion of the full
spacelike region of relative coordinates. We then established the energy
conditions on matter in order that the Einstein equations possess a
singularity in terms of $V(q),$ and showed that, for a class of power-law
attractive potentials, $V(q)\sim q^n,\;0<n\leq 3,$ the energy conditions
for a singularity to occur can be violated only in the case of local
$O(3,2)$ invariance of the $(x^\mu ,\tau )$ manifold.
We have found that an off-shell ensemble at high temperatures is
characterized by the equation of state $p=(\Gamma -1)\rho ;\;\;$ $p,\rho
\propto T^{\Gamma /(\Gamma -1)},$ with $\Gamma $ being equal to 3/2 in
the case of local $O(3,2)$ invariance of the $(x^\mu ,\tau )$
manifold\footnote{The denotement $\sigma $ stands for the 55-component of
the local metric on the $(x^\mu ,\tau )$ manifold which is $g^{\alpha
\beta }=(+,-,-,-,\sigma ).$}
($\sigma =1)$ and 5/4 in the case of local $O(4,1)$ invariance ($\sigma =
-1),$ so that in the latter case $p=\rho /4,\;\;$ $p,\rho \propto T^5.$
Off-shell matter\footnote{We use the term `off-shell' to describe matter
with continuous mass distribution, i.e., non-point spectrum, as for
off-shell states occuring in the propagators of quantum field theory.}
with the equation of state $p,\rho \propto T^5$ was introduced into the
standard cosmological model in ref. \cite{fried}. It was shown that such
matter has energy density comparable with that of standard radiation
(with the equation of state $p,\rho \propto T^4)$ at temperature $\sim
10^{12}$ K, so that the possibility for a phase transition from the
off-shell sector to the on-shell one (with possible compactification of
the fifth dimension \cite{N}), at critical temperature $\sim 10^{12}$ K,
should be taken into account; for example, in the case of a Bose gas, by
the mechanism of a high-temperature Bose-Einstein condensation
\cite{BEc}.
As we show in the present paper, a 5D cosmological model of Kaluza-Klein
type permits derivation of vacuum-, off-shell matter-, and
on-shell matter-dominated eras as the solutions of the corresponding 5D
gravitational field equations. These solutions enable one to construct
an inflationary scenario (inflationary solutions arise in a
vacuum-dominated era) according to which, as the universe expands and
cools down, a phase transition from the off-shell sector to the on-shell
one occurs, probably at temperature $\sim 10^{12}$ K \cite{fried}. We
study its effect on the rate of expansion and show that in both cases
of $(\sigma =-1)\rightarrow (\sigma =0)$ and $(\sigma =1)\rightarrow
(\sigma =0)$ phase transition\footnote{We use $\sigma =0$ to describe the
standard (on-shell) 3+1 case.} the expansion rate does not change.
We show that the model we are discussing does not expand adiabatically.
For the closed universe the thermodynamic entropy is a growing function
of cosmic time; for the flat and the open universe it can be a growing
function of historical time. The open and the closed models will be shown
to go to the 4D standard cosmological models as the universe expands, in
contrast to the flat model which does not have the corresponding limit.
We remark that some previous discussions of these questions have been
made in the framework of 5D Kaluza-Klein theory
\cite{MV}--\cite{PLW}. Mann and Vincent \cite{MV} have shown that the
vacuum (Kaluza-Klein type) solutons of the five-dimensional field
equations give rise to an effective radiation density ($\rho =3p)$
connected with the extra dimension. Ponce de Leon and Wesson \cite{PLW}
have interpreted the sourceless solutions of the five-dimensional
Kaluza-Klein equations as those of the four-dimensional Einstein
equations with effective matter properties. We also note that
inflationary models based on the Kaluza-Klein framework have been
considered by Shafi and Wetterich \cite{SW}, and by Gr{\o }n \cite{G}
who has derived a complete cosmological scenario within the framework
of Wesson's gravitational theory with the rest mass
as a fifth coordinate \cite{W}.
\section{The line element}
Similarly to \cite{therm}, we take the fifth-dimension subspace to be
homogeneous and without coupling to the other coordinate,
i.e., a maximally symmetric subspace of the 5D
space \cite{Wei}. Then the 5D metric becomes \cite{Wei}
\beq
^{(5)}ds^2=g_{\alpha \beta }dx^\alpha dx^\beta =g_{\mu \nu }(x^\rho )
dx^\mu dx^\nu +g_{55}(x^\rho )d\tau ^2,
\eeq
$$\alpha ,\beta =0,1,2,3,5;\;\;\;\mu ,\nu ,\rho =0,1,2,3.$$ As shown in
Appendix, the 5D gravitational field equations
\beq
^{(5)}R_{\alpha \beta }=8\pi G\left( ^{(5)}T_{\alpha \beta }-\frac{1}{3}
g_{\alpha \beta }\;^{(5)}T^\lambda _\lambda \right) ,
\eeq
with the source term
\beq
^{(5)}T_{\alpha \beta }=\left( \;^{(4)}T_{\mu \nu },\;p_5\right) ,\;\;\;
p_5=\sigma \mu _K\kappa ,
\eeq
where $\kappa $ is the density of the generalized Hamiltonian per unit
comoving three-volume (actually associated with the density of the
variable mass) and $\mu _K$ is the mass potential in relativistic
ensemble \cite{HSP}, reduce to the 4D Einstein equations
\beq
^{(4)}R_{\mu \nu }=8\pi G\left( ^{(4)}T_{\mu \nu }-\frac{1}{2}
g_{\mu \nu }\;^{(4)}T^\rho _\rho \right)
\eeq
in the case of no curvature in $\tau $ direction.
As an example of the use of the metric (2.1) in cosmology, consider a
spacially flat cosmological model with the line element (in refs.
\cite{MV,We,PLW} a similar structure has been used)
\beq
ds^2=e^{\bar{\nu }}dt^2-e^{\bar{\omega }}(dx^2+dy^2+dz^2)-e^{\bar{\mu }}
d\tau ^2,
\eeq
where $\bar{\nu },\bar{\omega },\bar{\mu }$ are assumed (here) to be
functions of time alone. A particular solution is obtained for
$\bar{\nu }=0,$ $e^{\bar{\omega }}=t,$
$e^{\bar{\mu }}=t^{-1}.$ In this case the line element
\beq
ds^2=dt^2-t(dx^2+dy^2+dz^2)-t^{-1}d\tau ^2
\eeq
is similar to the cosmological model found by Chodos and Detweiler
\cite{CD}. They interpreted the fifth dimension geometrically in the
usual Kaluza-Klein sense \cite{KK}. The time coordinate of the line
element (2.6) is the proper time shown on standard clocks at rest in the
3D spacial hyperplane orthogonal to the time- and $\tau $-directions.
This will in the following be referred to as ``cosmic time''. In the
case that 4D space-time is filled with a medium in which these clocks are
at rest, the coordinate system is said to be ``comoving''. These are the
usual terms from ordinary 4D cosmology. The expansion factor of the model
(2.6) is $R(t)=t^{1/2}.$ This universe expands too slowly to solve the
horizon and flatness problems \cite{Guth}.
By the proper choice of $s,$ the line element (2.5) can be reduced to
\beq
ds^2=dt^2-e^\omega (dx^2+dy^2+dz^2)-e^\mu d\tau ^2.
\eeq
For this line element, the nonvanishing Christoffel symbols are
(henceforth the prime stands for derivative with respect to the 5D line
element, and the dot for derivative with respect to cosmic time)
\beq
\Gamma ^5_{05}=\frac{\dot{\mu }}{2},\;\;\;\Gamma ^0_{55}=\frac{1}{2}
\dot{(e^\mu )}.
\eeq
Therefore, the geodesic equations for $t$ and $\tau $ read (see Appendix)
\beq
t^{''}+\frac{1}{2}\dot{(e^\mu )}(\tau ^{'})^2=0,
\eeq
\beq
\tau ^{''}+\dot{\mu }t^{'}\tau ^{'}=0.
\eeq
The Lagrangian of a free comoving particle is
\beq
L=e^\mu (\tau ^{'})^2-(t^{'})^2.
\eeq
Since $\tau $ is a cyclic coordinate (it does not appear in the
Lagrangian), the conjugate momentum
\beq
p_\tau =\frac{\partial L}{\partial \tau ^{'}}=e^\mu \tau ^{'}
\eeq
is a constant of motion, giving
\beq
\frac{d\tau }{ds}=p_\tau e^{-\mu }.
\eeq
Inserting (2.13) into Eq. (2.9) gives
\beq
t^{''}=-\frac{p_\tau ^2}{2}e^{-\mu }\dot{\mu }=\frac{p_\tau ^2}{2}(e^{
-\mu })^{'}(t^{'})^{-1}.
\eeq
Integration leads to
\beq
\frac{ds}{dt}=\left( p_\tau ^2e^{-\mu }+C^2\right) ^{-1/2},
\eeq
where $C$ is an arbitrary constant. It then follows from (2.13),(2.15)
that
\beq
\frac{d\tau }{dt}=\frac{p_\tau e^{-\mu }}{(p_\tau ^2e^{-\mu }+C^2)^{1/2}
}.
\eeq
We now shall consider the following generalization of the line element
(2.5) which lets a spacial curvature be different from zero and permits
direct comparison with the Kaluza-Klein models\footnote{The 5D metric
(2.18) corresponds to local $O(4,1)$ invariance of an extended $(x^\mu ,
\tau )$ manifold. The choice of the metric in the form
\beq
ds^2=dt^2-\frac{R^2(t)}{(1+\frac{1}{4}kr^2)^2}(dx^2+dy^2+dz^2)+A^2(t)d
\tau ^2,
\eeq
corresponding to local $O(3,2)$ invariance of an $(x^\mu ,\tau )$
manifold, will lead essentially to the results of this work. Both cases
(2.17),(2.18) are treated simultaneously in the system of the field
equations (3.2)-(3.4).}:
\beq
ds^2=dt^2-\frac{R^2(t)}{(1+\frac{1}{4}kr^2)^2}(dx^2+dy^2+dz^2)-A^2(t)d
\tau ^2,
\eeq
where $r^2=x^2+y^2+z^2,$ and $k=0,\pm 1$ characterizes the spacial
curvature. Comparison with Eq. (2.7) shows that $e^\mu =A^2,$ so that,
expressed in terms of $A,$ Eqs. (2.13), (2.16) take on the form
\beq
\frac{ds}{d\tau }=\frac{A^2}{p_\tau },
\eeq
\beq
\frac{ds}{dt}=(p_\tau ^2/A^2+C^2)^{-1/2},
\eeq
\beq
\frac{dt}{d\tau }=\frac{(p_\tau ^2/A^2+C^2)^{1/2}}{p_\tau /A^2}.
\eeq
Since the standard case of no curvature in $\tau $ direction corresponds
to $A={\rm const},$ in this case $ds/d\tau ={\rm const},$ so that the
line element reduces (as seen in Eq. (2.17),(2.18); see also Appendix) to
\beq
d\tau ^2=dt^2-\frac{R^2(t)}{(1+\frac{1}{4}kr^2)^2}(dx^2+dy^2+dz^2),
\eeq
which in turn can be reduced to the standard 4D Robertson-Walker metric
\beq
d\tau ^2=dt^2-R^2(t)\left[ \frac{d\rho ^2}{1-k\rho ^2}+\rho ^2(d\theta
^2+\sin ^2\theta d\phi ^2)\right]
\eeq
with the help of the transformation \cite{Rin}
\beq
\rho =\frac{r}{1+\frac{1}{4}kr^2}.
\eeq
We see that, as $A\rightarrow {\rm const},$ the 5D universe with the line
element (2.17),(2.18) passes over to the standard 4D Robertson-Walker
universe\footnote{We remark that the universe with the Robertson-Walker
type metric $$ds^2=d\tau ^2-R^2(\tau )\left[ \frac{d\rho ^2}{1-k\rho ^2}+
\rho ^2\Big( d\theta ^2+\sin ^2\theta (-d\beta ^2+\cosh ^2\beta d\phi
^2)\Big)\right], $$
where $\rho ,\theta ,\beta ,\phi $ are the coordinates in the restricted
Minkowski space (RMS) \cite{therm}, as $R(\tau )\rightarrow \infty ,$
passes over to the standard 4D Robertson-Walker universe as well.
Details will be explained elsewhere \cite{RMS}.} (a phase transition to
the on-shell sector at $T\sim 10^{12}$ K probably taking place).
\section{The field equations}
We consider the following field equations for a 5D space-time filled with
a perfect fluid, permitting a non-vanishing cosmological constant
(henceforth we shall use the system of units in which $c=8\pi G=1):$
\beq
R_{\alpha \beta }-\frac{1}{2}g_{\alpha \beta }R+\Lambda g_{\alpha \beta }
=T_{\alpha \beta },
\eeq
with the source term (2.3). We note that the cosmological constant could
be disposed of by considering instead space-time with a vacuum fluid
which is a perfect fluid with the equation of state $p=-\rho .$
The field equations (3.1) with $T_{\alpha \beta }$ in the form (2.3)
reduce to the following system \cite{AG,MV}:
\beq
\frac{\dot{R}^2+k}{R^2}+\frac{\dot{R}\dot{A}}{RA}=\frac{1}{3}(\Lambda +
\rho ),
\eeq
\beq
\frac{2\ddot{R}}{R}+\frac{\dot{R}^2+k}{R^2}+\frac{2\dot{R}\dot{A}}{RA}+
\frac{\ddot{A}}{A}=\Lambda -p,
\eeq
\beq
\frac{\ddot{R}}{R}+\frac{\dot{R}^2+k}{R^2}=\frac{1}{3}(\Lambda +p_5).
\eeq
The usual 4D equations of the Friedmann model are obtained by setting $A=
{\rm const}$ in Eqs. (3.2),(3.3) and neglecting Eq. (3.4).
The energy-momentum conservation $T^{\alpha \beta }_{;\beta }=0$ implies
\cite{AG,MV}
\beq
\dot{\rho }+3(\rho +p)\frac{\dot{R}}{R}+(\rho -p_5)\frac{\dot{A}}{A}=0.
\eeq
For the initial stage of the evolution, when the universe is hot, we can
use \cite{therm}
\beq
p_5=\sigma p.
\eeq
This result, in fact, follows easily from the definition of the
five-dimensional energy-momentum tensor (also discussed in \cite{therm})
which is obtained by
the extension of the usual energy-momentum tensor
\beq
T^{\mu \nu }=(p+\rho )u^\mu u^\nu -pg^{\mu \nu },\;\;\;u^\rho u_\rho =1
\eeq
to a five-dimensional form:
\beq
^{(5)}T_{\alpha \beta }=\left(\;^{(5)}T_{\mu \nu },\;^{(5)}T_{55}\right);
\eeq
the requirement that the limiting case of the corresponding gravitational
theory (for zero curvature in the $\tau $ direction) coincides with the
Einstein equations results in the identification (see Appendix) $^{(5)}T_
{\mu \nu }=\;^{(4)}T_{\mu \nu }$ and $^{(5)}T_{55}=\sigma \mu _K\kappa. $
Expressions for $p$ and $\rho ,$ using the grand canonical
ensemble obtained by Horwitz, Schieve and Piron \cite{HSP} in their study
of manifestly covariant statistical mechanics, were found in \cite{ind}
in terms of confluent hypergeometric functions. For $T$ small, one finds
that $p,\rho \propto T^6,$ $\rho \simeq 5p,$ and, in fact, that $\mu _K
\kappa \propto T^7$ is negligible in comparison with $\rho .$ On the
other hand, for $T$ large, one finds \cite{therm} that in the case of
local $O(4,1)$ invariance of $(x^\mu ,\tau )$ manifold, $p,\rho \propto
T^5,$ $p\simeq \rho /4\simeq \mu _K\kappa ,$ while in the case of local
$O(3,2)$ invariance, $p,\rho \propto T^3,$ $p \simeq \rho /2\simeq
\mu _K\kappa .$ For high
temperature, it therefore follows that (as discussed in \cite{therm})
\beq
T^{\alpha \beta }=(p+\rho )u^\alpha u^\beta -pg^{\alpha \beta },\;\;\;
u^\lambda u_\lambda =1,
\eeq
so that, in the local rest frame,
$T_{\alpha \beta }={\rm diag}\;(\rho ,-p,-p,-p,\sigma p),$ and Eq. (3.6)
is justified. Moreover, for a perfect fluid, we use the equation of state
\beq
p=(\Gamma -1)\rho ,
\eeq
where $\Gamma $ is a constant. It then follows \cite{therm} that in the
cases of $\sigma =-1$ $(O(4,1))$ and $\sigma =1$ $(O(3,2)),$ $\Gamma $
is equal to 5/4 and 3/2, respectively, so that the tensor (3.9) is
traceless in either case. We note that the form (3.9) of a source term
for the 5D field equations has been used by Wesson \cite{W}.
In the case of local $O(4,1)$ invariance of an $(x^\mu ,\tau )$ manifold,
$\sigma =-1,$ $p_5=-p,$ and Eq. (3.5) integrates to
\beq
R^{3\Gamma }A^{2-\Gamma }\rho ={\rm const}.
\eeq
The expansion is not adiabatic. As $A\rightarrow {\rm const,}$ Eq. (3.11)
takes on the standard form
\beq
R^{3\Gamma }\rho ={\rm const}.
\eeq
Similarly, in the case of local $O(3,2)$ invariance of an $(x^\mu ,
\tau )$ manifold, $\sigma =1,$ $p_5=p,$ and Eq. (3.5) gives
\beq
R^{3\Gamma }A^\Gamma \rho ={\rm const},
\eeq
wich again reduces to the standard form (3.12) as $A\rightarrow {\rm
const}.$ Note that both Eqs. (3.11) and (3.13) give for dust matter with
$p\approx 0$ $(\Gamma =1)$ (this can also be obtained
directly from (3.5) with $p=p_5=0)$
\beq
R^3A\rho ={\rm const},
\eeq
which, as $A\rightarrow {\rm const},$ reduce to the standard result
\beq
R^3\rho ={\rm const}.
\eeq
\section{Solutions to the field equations}
The standard strategy for solving cosmological equations, i.e., to
exclude $\rho $ and $p$ from the equation for $R$ in terms of $\rho ,p,$
with the help of the equations of state and energy-momentum conservation:
$p=(\Gamma -1)\rho ,$ $\rho \propto R^{-n},\;n=3,4,$ does not work in our
case, since, in view of (3.11),(3.13), $\rho \propto R^{-p}A^{-q}$
and $A$ is a function of cosmic time. We therefore have to express $R$ in
terms of a parameter which is independent of $A.$ Such a parameter is a
cosmological constant $\Lambda .$
We shall suppose that $A$ is a slowly varying function of $t,$ so that
one can neglect the term $\ddot{A}/A$ in Eq. (3.3). Then, for $\sigma =
-1$ $(O(4,1)),$ one derives from (3.2)-(3.4), by the exclusion of $\rho ,
p$ and $p_5$ with the help of (3.6),(3.10), the equation
\beq
\frac{\Gamma +1}{2(2\Gamma -1)}\ddot{R}R+\dot{R}^2+k=\frac{\Gamma }{3(2
\Gamma -1)}\Lambda R^2,
\eeq
which for $\Gamma =5/4$ reduces to
\beq
\frac{3}{4}\ddot{R}R+\dot{R}^2+k=\frac{5}{18}\Lambda R^2.
\eeq
For $\sigma =1$ $(O(3,2))$ one similarly obtains
\beq
\frac{5\Gamma -3}{2\Gamma }\ddot{R}R+\dot{R}^2+k=\frac{2\Gamma -1}{3
\Gamma }\Lambda R^2,
\eeq
which for $\Gamma =3/2$ reduces to
\beq
\frac{3}{2}\ddot{R}R+\dot{R}^2+k=\frac{4}{9}\Lambda R^2.
\eeq
For the standard case of $p_5=0$ (or $\sigma =0)$ one gets
\beq
\ddot{R}R+\dot{R}^2+k=\frac{1}{3}\Lambda R^2,
\eeq
which really represents Eq. (3.4) with $p_5=0.$ Note that both Eqs. (4.1)
and (4.3) reduce to (4.5) for $\Gamma =1.$
\subsection{Vacuum-dominated era}
In a vacuum-dominated era the universe is filled with a vacuum fluid.
Eqs. (4.2),(4.4), \\ (4.5) can be represented by an equation of the
general form
\beq
a\ddot{R}R+\dot{R}^2+k=b\Lambda R^2,
\eeq
which, through the substitution
\beq
R^{1+\frac{1}{a}}=\tilde{R},
\eeq
reduces to the equation
\beq
\ddot{\tilde{R}}-\frac{b(a+1)}{a^2}\Lambda \tilde{R}+\frac{a+1}{a^2}k
\tilde{R}^{\frac{1-a}{1+a}}=0
\eeq
having the solution ($C_1,C_2={\rm const})$
\beq
\tilde{R}=C_1\cosh \sqrt{\frac{b(a+1)}{a^2}\Lambda }\;t+C_2\sinh \sqrt{
\frac{b(a+1)}{a^2}\Lambda }\;t+\left( \frac{k}{b\Lambda }\right) ^{\frac{
1}{2}(1+\frac{1}{a})}.
\eeq
We, therefore, obtain from (4.2),(4.4),(4.5), respectively: \\
for $\sigma =-1,$
\beq
R^{7/3}=\left( \frac{18}{5}\frac{k}{\Lambda }\right) ^{7/6}+C_1^{-}\cosh
\frac{\sqrt{70\Lambda }}{9}t+C_2^{-}\sinh \frac{\sqrt{70\Lambda }}{9}t,
\eeq
for $\sigma =1,$
\beq
R^{5/3}=\left( \frac{9}{4}\frac{k}{\Lambda }\right) ^{5/6}+C_1^{+}\cosh
\frac{\sqrt{40\Lambda }}{9}t+C_2^{+}\sinh \frac{\sqrt{40\Lambda }}{9}t,
\eeq
for $\sigma =0,$
\beq
R^2=\frac{3k}{\Lambda }+C_1^{0}\cosh \frac{\sqrt{54\Lambda }}{9}t+
C_2^{0}\sinh \frac{\sqrt{54\Lambda }}{9}t.
\eeq
We note that the solution (4.12) was obtained previously by Gr{\o }n
\cite{G}. In subsequent consideration we shall, for simplicity, restrict
ourselves to this solution alone. It then follows that, without any loss
of generality, this solution can be represented
by the following relations:
$k=1,$
\beq
R^2=\frac{2}{\omega ^2}(1+\cosh \omega t)=\frac{4}{\omega ^2}\cosh ^2
\frac{\omega t}{2},
\eeq
$k=-1,$
\beq
R^2=\frac{2}{\omega ^2}(\cosh \omega t-1)=\frac{4}{\omega ^2}\sinh ^2
\frac{\omega t}{2},
\eeq
$k=0,$
\beq
R^2=\frac{4}{\omega ^2}\exp\;(\omega t),
\;\;\;w\equiv \sqrt{\frac{2}{3}\Lambda }.
\eeq
Consider, for example, (4.13). For $t=0$ it gives $R=2/\omega .$ Since
the classical description of the expansion of the universe cannot be
valid prior to $t\sim t_{Pl}=M_{Pl}^{-1}\sim 5\cdot 10^{-44}$ s after the
big bang or the start of inflation at $t=0,$ one finds that $\omega
\stackrel{<}{\sim }M_{Pl}\sim 10^{19}$ GeV, and therefore $\Lambda =
\frac{3}{2}\omega ^2\stackrel{<}{\sim }10^{38}$ GeV$^2.$ By introducing
the vacuum energy density through the relation
\beq
\Lambda c^2=8\pi G\rho _{vac}
\eeq
and recovering $c$ and $G$ for numerical calculation, $G\sim M_{Pl}^{-2}
\simeq 10^{-38}$ GeV$^{-2},$ one obtains
\beq
\rho _{vac}<M_{Pl}^4/16\sim 10^{75}\;{\rm GeV}^4\simeq 10^{92}\;
{\rm g\;cm}^{-3}.
\eeq
If, similar to the standard inflationary models \cite{Linde}, one takes
$\rho _{vac}=T_c^4\sim 10^{60}$ GeV$^4,$ where $T_c\sim 10^{15}$ GeV is
a typical critical temperature for a phase transition in grand unified
theories \cite{Linde}, one obtains $\omega =\sqrt{16\pi G\rho _{vac}/3c^
2}\simeq 4\cdot 10^{11}$ GeV. As is usually done in the standard
inflationary models \cite{Linde}, inflation comes to an end when its rate
$H\equiv \dot{R}/R=\omega /2$ begins to decrease rapidly (which means
that the universe becomes rapidly increasing in size), the typical
time of inflation is $t_{inf}\sim 1/H=2/\omega .$ With $\omega \simeq 4
\cdot 10^{11}$ GeV, one finds $t_{inf}\simeq 10^{-36}$ s.
The value of the vacuum energy density (4.16) should be related to the
present-day vacuum energy density which is not much greater in absolute
value than the critical density $\rho _{cr}\sim 10^{-29}\;{\rm g\;cm}^{
-3},$ as implied by recent cosmological data. As remarked by Linde, in
grand unified theories (e.g., in the $SU(5)$ Coleman-Weinberg theory
\cite{CW}) this value of the vacuum energy density is obtained as a
result of a series of phase transitions. Other theories having the
cosmological constant (and, therefore, vacuum energy density) decreasing
with time are also discussed in the literature (e.g., a scale-covariant
theory of fundamental interactions \cite{Wes} in which $\Lambda \propto
t^{-2}).$ In ref. \cite{RSS} in which $A(t)$ is related to the quantum
one-loop correction terms as $\rho \sim -p_5\sim A^{-5},$ the problem is
surmounted by demanding that $A(t)$ be constant with a value cancelling
out the contribution from the 5D cosmological constant, producing a zero
effective 4D one.
As usually done in standard inflationary models \cite{Linde}, when
inflation ends, the cosmological constant $\Lambda $ is omitted in the
field equations (3.2)-(3.4), and subsequent evolution is described by a
Friedmann-type hot universe model. One may also think that the
cosmological constant is contained (through vacuum energy density) as
part of the off-shell matter energy density. This may have a reasonable
basis, since one notes that the off-shell matter energy density with
temperature dependence $\sim T^5$ \cite{ind} which at $T\sim 10^{28}$ K
$(=10^{15}$ GeV) is equal to $10^{75}$ GeV$^4$ can be consistently
represented by $$\rho ^{'}=10^{75}\left( \frac{T}{10^{28}}\right) ^5{\rm
GeV}^4;$$ it then follows from this formula that $\rho ^{'}$ takes on the
value $10^{-4}\;{\rm GeV}^4\sim 10^{14}\;{\rm g\;cm}^{-3},$ which is a
typical energy density of radiation-like matter at $T\sim 2\cdot 10^{12}
\;{\rm K}\simeq 150\;{\rm MeV},$ at the same temperature, implying the
possibility of a phase transition. Such a phase transition is briefly
analyzed in Section 5.
\subsection{Off-shell matter-dominated era}
In an off-shell matter-dominated era the universe is filled with an
off-shell fluid having the equation of state (3.10) with $$\Gamma =
\left\{ \begin{array}{ll}
5/4, & \sigma =-1, \\
3/2, & \sigma =1.
\end{array} \right. $$ In this case, as discussed below, we omit
$\Lambda $ in the field equations (3.2)-(3.4). Moreover, we can also omit
the spacial curvature $k$ which is negligible at high energy densities.
It then follows from Eq. (4.6) with zero r.h.s.
(this also follows from (4.9) for small $\sqrt{\Lambda }t)$ that
\beq
\tilde{R}=C_1^{'}+C_2^{'}t,
\eeq
so that
for $\sigma =-1,$
\beq
R^{7/3}=C_1^{'-}+C_2^{'-}t,
\eeq
for $\sigma =1,$
\beq
R^{5/3}=C_1^{'+}+C_2^{'+}t.
\eeq
\subsection{On-shell matter-dominated era}
In an on-shell matter-dominated era the universe is filled with a
standard on-shell radiation having the equation of state $p=1/3\rho ,$
and, as the universe expands and cools down, with dust matter with $p
\approx 0.$
For the universe filled with radiation, we omit both $\Lambda $ and
$k$ in the corresponding equation (4.5), which then has the solution
\beq
R^2=C_1^{'0}+C_2^{'0}t.
\eeq
Since Eqs. (4.19),(4.20) and (4.21) are obtained from (3.2)-(3.4) with
$p_5=\sigma p,$ and $p_5=0,$ respectively, the three Eqs. (4.19)-(4.21)
can be unified in one equation, as follows:
\beq
R^{2-\sigma \alpha /3}=C_1^{'\sigma }+C_2^{'\sigma }t,
\eeq
where
$$\alpha =\left\{ \begin{array}{ll}
1, & T\rightarrow \infty , \\
0, & T\rightarrow \;0,
\end{array} \right. $$ according to (see Eqs. (A.9),(A.11) of Appendix)
$$p_5=\left\{ \begin{array}{ll}
\sigma p, & T\rightarrow \infty , \\
0, & T\rightarrow \;0.
\end{array} \right. $$
For the universe filled with dust matter, we omit $\Lambda $ alone
in Eq. (4.5); this equation with zero r.h.s. has the solution
\beq
R^{2}=C_1^{''0}+C_2^{''0}t-kt^2.
\eeq
Two possible scenarios of evolution of the universe described by these
equations exist.
First, $\alpha $ in Eq. (4.22) is a smooth function of $T.$ As the
universe expands and its temperature decreases, $\alpha \rightarrow 0,$
so that in both cases $(\sigma =\pm 1)$ Eq. (4.22) passes over smoothly
to Eq. (4.21) which goes over to Eq. (4.23) at lower temperatures. That
is, the universe passes smoothly from an off-shell matter-dominated era
to on-shell one. In this case the rate of expansion is a smooth function
of temperature (and therefore the radius of the universe) as well.
Second, $\alpha $ is not a smooth function of $T,$ nor may it be a
function of $T$ at all. Since at some value of $R$ (and therefore $T)$
Eq. (4.19) (or (4.20)) goes over to Eq. (4.21), and the powers of $R$ in
the corresponding equations do not coincide, passage from an off-shell
matter-dominated era to on-shell one occurs as a {\it phase transition.}
In this case the rate of expansion does not change in either case of
$\sigma =-1$ or $\sigma =1,$ as we shall see below).
We consider the second scenario to be more realistic one, since a
passage from the off-shell sector (a sector of relativistic mass
distributions \cite{BH1,ind}) to on-shell one is probably a
phase transition\footnote{Another possibility is a smooth Galilean limit
$c\rightarrow \infty $ \cite{BH2}.}, as discussed in ref.
\cite{BEc} in the case of a relativistic Bose gas.
We now wish to discuss this phase transition in general features.
\section{Phase transition from off-shell matter-dominated era to on-shell
one}
Using Eqs. (4.19)-(4.21), we obtain the relations representing
continuity of $R$ at $t_0,$ where $t_0$ is the moment of cosmic time at
which the phase transition occurs:
$(\sigma =-1)\rightarrow (\sigma =0),$
\beq
(C_1^{'-}+C_2^{'-}t_0)^{3/7}=(C_1^{'0}+C_2^{'0}t_0)^{1/2},
\eeq
$(\sigma =1)\rightarrow (\sigma =0),$
\beq
(C_1^{'+}+C_2^{'+}t_0)^{3/5}=(C_1^{'0}+C_2^{'0}t_0)^{1/2}.
\eeq
We shall study the effect of the phase transition on the expansion rate.
We restrict our consideration to the case where the phase transition
occurs smoothly and adiabatically\footnote{More general cases of a
cosmological phase transition are considered in ref. \cite{GK} on the
example of a hadronic matter--the quark-gluon plasma phase transition.}.
It follows from Eqs. (3.2)-(3.4),(3.6),(3.10), through the exclusion of
$\Lambda ,$ that the following equations for $R$ in terms of $\rho $
hold,
\beq
\dot{R}^2=\frac{1}{2}\ddot{R}R+\rho \left[ \frac{\Gamma -1}{2}\left( 1+
\frac{\sigma }{3}\right) +\frac{1}{3}\right] -k,
\eeq
which reduces, in the corresponding cases, to:
for $\sigma =-1,\;\Gamma =5/4,$
\beq
\dot {R}^2=\frac{1}{2}\ddot{R}R+\frac{5}{12}\rho R^2-k,
\eeq
for $\sigma =1,\;\Gamma =3/2,$
\beq
\dot {R}^2=\frac{1}{2}\ddot{R}R+\frac{2}{3}\rho R^2-k,
\eeq
for $\sigma =0,\;\Gamma =4/3$ (standard case),
\beq
\dot {R}^2=\frac{1}{2}\ddot{R}R+\frac{1}{2}\rho R^2-k.
\eeq
A smooth transition occurs at the constant pressure.
Using the corresponding equations of state $\rho =p/(\Gamma -1)$ and
Eqs. (5.4)-(5.6) (in which we neglect $k),$ we find the following
relations which represent equality of pressure in the corresponding
phases (one simply equates the quantities $(\Gamma -1)\rho R^2(t_0)):$
$(\sigma =-1)\rightarrow (\sigma =0),$
\beq
\frac{9}{49}\left( C_2^{'-}\right) ^2(C_1^{'-}+C_2^{'-}t_0)^{-8/7}=
\frac{1}{4}\left( C_2^{'0}\right) ^2(C_1^{'0}+C_2^{'0}t_0)^{-1},
\eeq
$(\sigma =1)\rightarrow (\sigma =0),$
\beq
\frac{9}{25}\left( C_2^{'+}\right) ^2(C_1^{'+}+C_2^{'+}t_0)^{-4/5}=
\frac{1}{4}\left( C_2^{'0}\right) ^2(C_1^{'0}+C_2^{'0}t_0)^{-1}.
\eeq
Calculation of $\dot {R}(t_0)$ gives, respectively,
for $\sigma =-1,$
\beq
\frac{3}{7}C_2^{'-}(C_1^{'-}+C_2^{'-}t_0)^{-4/7},
\eeq
for $\sigma =1,$
\beq
\frac{3}{5}C_2^{'+}(C_1^{'+}+C_2^{'+}t_0)^{-2/5},
\eeq
for $\sigma =0,$
\beq
\frac{1}{2}C_2^{'0}(C_1^{'0}+C_2^{'0}t_0)^{-1/2}.
\eeq
Comparison of Eqs. (5.9),(5.10) (squared) with Eq. (5.11) (squared),
using Eqs. (5.7) and (5.8), shows that $\dot {R}(t_0)$ in both $\sigma =-
1$ and $\sigma =1$ phases coincide with $\dot{R}(t_0)$ in the $\sigma =0$
phase; since $R(t_0)$ is the same for the three, we conclude that the
rate of expansion,
\beq
H\equiv \frac{\dot {R}}{R},
\eeq
{\it does not change} in either case of the $(\sigma =-1)\rightarrow
(\sigma =0)$ or $(\sigma =1)\rightarrow (\sigma =0)$ phase transitions.
This observation suggests that the phase transition should be
sufficiently smooth (second order). Although a first order phase
transition might be preferable for cosmological implications, due to the
fluctuations which are generated at the transition\footnote{The
fluctuations could not directly affect galaxy formation, since the
horizon size at the time of the transition is on a planetary scale
\cite{hor}. It has been demonstrated \cite{CS} that they could produce
planetary mass black holes; these black holes could provide a possible
explanation for the dark matter of the universe and even be seeds in
galaxy formation \cite{FPS,Wi}.}, experimental indications on the order
of this phase transition are still absent. Indeed, cosmological phase
transition at $T_c\sim 150$ MeV is normally associated with the
transition from a strongly interacting hadronic phase to a weakly
interacting quark-gluon plasma phase \cite{GK,OW}. Presently
available lattice data on $SU(N)$ pure gauge theory lattice simulations
indicate that a phase transition to a weakly interacting phase is of
apparently first order for $SU(3)$ and second order for $SU(2)$ theory
\cite{Mul}. In ref. \cite{Br}, however, it is argued that the apparent
first order nature of the transition in the case of $SU(3)$ pure gauge
theory may well be a lattice artefact. Moreover, there are indications
from lattice QCD calculations that when fermions are included, the phase
transition may be of second or higher order \cite{Ben}. In this case, as
remarked by Ornik and Weiner \cite{OW}, the phase transition would be
hardly distinguishable from a situation in which no phase transition
would have taken place (radiation-dominated universe alone).
\section{``Generalized'' entropy and the behavior of $A$ in an on-shell
matter-dominated era}
As we have seen in Section 3, the expansion of the universe with the line
element (2.17) is not adiabatic, due to the presence of time-dependent
$A$ in the equation (3.5) for energy-momentum conservation. Rewriting
this equation in the form
\beq
\left[ R^3A\rho \right] ^\bullet +Ap\dot {R^3}-p_5R^3\dot{A}=0,
\eeq
we see that the universe can be characterized by the ``first law''
\beq
d(AE)=ATd\tilde{S}-ApdV+p_5VdA,
\eeq
so that, in view of (6.1),(6.2) and $V\sim R^3,$
the ``generalized'' entropy $\tilde{S}$ is conserved:
\beq
\frac{d\tilde{S}}{dt}=0.
\eeq
It follows from (6.2) and genuine first law
\beq
dE=TdS-pdV
\eeq
that the thermodynamic entropy are related to the ``generalized'' one
as follows:
\beq
dS=d\tilde{S}-\frac{E^{'}}{T}\frac{dA}{A},
\eeq
where $E^{'}\equiv E-p_5V=\rho ^{'}V,$ and $$\rho ^{'}\equiv \rho -p_5$$
is the ``reduced'' energy density \cite{therm}. Hence, in view of (6.3),
\beq
\frac{dS}{dt}=-\frac{E^{'}}{T}\frac{1}{A}\frac{dA}{dt}.
\eeq
One sees that the sign of $\frac{dS}{dt}$ is determined by the sign of
$-\frac{1}{A}\frac{dA}{dt}.$ Note that, as $A\rightarrow {\rm const},$
it follows from (6.5) that $S=\tilde{S}+{\rm const}.$
In general, the time dependence of $A$ can be derived from Eqs.
(3.2)-(3.4), (3.6), (3.10), provided that the corresponding time
dependence of $R$ is known. We shall restrict ourselves to the on-shell
matter-dominated universe (similar consideration for the vacuum- and the
off-shell matter-dominated eras does not seem to present a difficulty).
For the universe filled with radiation-like matter, Eqs. (3.2)-(3.4) take
on the form
\beq
\frac{\dot{R}^2}{R^2}+\frac{\dot{R}\dot{A}}{RA}=\frac{1}{3}\rho =p,
\eeq
\beq
\frac{2\ddot{R}}{R}+\frac{\dot{R}^2}{R^2}+\frac{2\dot{R}\dot{A}}{RA}+
\frac{\ddot{A}}{A}=-p,
\eeq
\beq
\frac{\ddot{R}}{R}+\frac{\dot{R}^2}{R^2}=0.
\eeq
It follows from Eq. (6.9) that $$R^2=C^{'0}_1+C^{'0}_2t,$$ in agreement
with (4.21). Summing up Eqs. (6.7) and (6.8), taking into account (6.9),
gives
\beq
\frac{3\dot{R}\dot{A}}{RA}+\frac{\ddot{A}}{A}=0.
\eeq
This equation has the solution
\beq
\dot{A}=\frac{C}{R^3},\;\;\;C={\rm const};
\eeq
hence
\bqry
A & = & C\int \frac{dt}{R^3}\;=\;C\int \frac{dt}{(C^{'0}_1+C^{'0}_2t)^
{3/2}} \NL
& = & \frac{-2C/C^{'0}_2}{(C^{'0}_1+C^{'0}_2t)^{1/2}}+\beta \;=\;
\frac{\alpha }{R}+\beta ,\;\;\;\alpha ,\beta ={\rm const}.
\eqry
Therefore
\beq
-\frac{1}{A}\frac{dA}{dt}=\frac{\dot{R}}{R(1+\frac{\beta }{\alpha }R)}.
\eeq
One sees that, since $\dot{R}>0,$ if $\alpha $ and $\beta $ are of the
same sign, or if $\beta =0,$ then $-\frac{1}{A}\frac{dA}{dt}>0,$ and
therefore $\frac{dS}{dt}>0,$ in view of (6.6).
For the dust universe, it follows from Eqs. (3.3),(3.4) with zero r.h.s.
that
\beq
\frac{\ddot{R}}{R}+\frac{2\dot{R}\dot{A}}{RA}+\frac{\ddot{A}}{A}=0,
\eeq
which reduces to the equation
\beq
\ddot{(AR)}=0,
\eeq
which has the solution (taking into account (4.23))
\beq
A=\frac{a+bt}{R}=\frac{a+bt}{\sqrt{C^{''0}_1+C^{''0}_2t-kt^2}},\;\;\;
a,b={\rm const}.
\eeq
The substitution
\beq
\rho =\frac{3\gamma }{AR^3},\;\;\gamma ={\rm const}
\eeq
(which follows from (3.14)) in the r.h.s. of Eq. (3.2) (in which we
omit $\Lambda ,$ as usual) yields, with the help of Eq. (3.4) with
zero r.h.s.,
\beq
\dot{R}\dot{A}-A\ddot{R}=\frac{\gamma }{R^2}.
\eeq
One can then find that Eqs. (6.16) and (6.18) are compatible if
\beq
\left\{ \begin{array}{ccc}
k & = & 0, \\
bC^{''0}_2 & = & 2\gamma ,
\end{array} \right.
\eeq
\beq
\left\{ \begin{array}{ccc}
k & \neq & 0, \\
a & = & \gamma k, \\
bC^{''0}_2 & = & 0,
\end{array} \right.
\eeq
and the same relations (6.19),(6.20) with $C^{''0}_1=0.$ Moreover, the
solutions (4.23) for $R$ and (6.16) for $A$ should be matched\footnote{By
matching we mean continuity of a function and its first derivative.} with
the corresponding solutions (4.21),(6.12) for the radiation-like
universe, which we rewrite here:
\beq
\left\{ \begin{array}{ccl}
R & = & \sqrt{C^{'0}_1+C^{'0}_2t}, \\
A & = & \frac{\alpha }{\sqrt{C^{'0}_1+C^{'0}_2t}}+\beta .
\end{array} \right.
\eeq
Thus, one is left with the following solutions for the dust universe:
for $k=0,$
\beq
\left\{ \begin{array}{ccl}
R & = & \sqrt{C^{''0}_1+C^{''0}_2t}, \\
A & = & \frac{a+2\gamma t/C^{''0}_2}{\sqrt{C^{''0}_1+C^{''0}_2t}},
\end{array} \right.
\eeq
for $k=1,$
\beq
\left\{ \begin{array}{ccl}
R & = & \sqrt{C^{''0}_1+C^{''0}_2t-t^2}, \\
A & = & \frac{\gamma }{\sqrt{C^{''0}_1+C^{''0}_2t-t^2}},
\end{array} \right.
\eeq
for $k=-1,$
\beq
\left\{ \begin{array}{ccl}
R & = & \sqrt{C^{''0}_1+t^2}, \\
A & = & \frac{bt-\gamma }{\sqrt{C^{''0}_1+t^2}}.
\end{array} \right.
\eeq
\subsection{Flat dust universe}
Consider first the case of the flat dust universe. It follows from (6.22)
that, as $t\rightarrow \infty ,$
\beq
A\sim R\sim t^{1/2}.
\eeq
Therefore, $-\frac{1}{A}\frac{dA}{dt}=-\frac{1}{R}\frac{dR}{dt}=-H(t)<0,$
so that the entropy is a decreasing function of cosmic time. In this
case, as seen in Eq. (2.21), if $p_\tau <0,$
\beq
\frac{dS}{d\tau }=\frac{dt}{d\tau }\frac{dS}{dt}>0,
\eeq
i.e., the entropy is a growing function of {\it historical
time}; if we moreover take $\alpha $ and $\beta $ in (6.12) of the
opposite sign, we will have, in view of (6.13), $\frac{dS}{d\tau }>0$
during all the on-shell matter-dominated era.
It is seen in Eq. (6.25) that for the model (6.22), the limit
$A\rightarrow {\rm const}$ is absent, i.e., it does not go over to the
standard 4D cosmological flat model, but rather represents the model
(2.5) with $e^{\bar{\omega }}=e^{\bar{\mu }}=t.$
\subsection{Closed dust universe}
Now we turn to the case of the closed dust universe. As seen in Eqs.
(6.23), the universe expands until the moment
\beq
t_0=\frac{C^{''0}_2}{2},
\eeq
reaching the maximal radius, and then begins to contract. For this model,
\beq
\frac{dA}{dt}=\frac{\gamma (t-\frac{C^{''0}_2}{2})}{(C^{''0}_1+C^{''0}_
2t-t^2)^{3/2}}=\frac{\gamma (t-\frac{C^{''0}_2}{2})}{R^3},
\eeq
so that, as $t\rightarrow t_0,$ $\frac{dA}{dt}\simeq 0,$ i.e., $A\simeq
{\rm const}.$ Therefore, in this limit the line element (2.17) goes over
to the standard 4D one, (2.23), and the 5D gravitational field equations
become indistinguishable from the standard 4D Einstein equations.
Since $A=\gamma /R,$
\beq
-\frac{1}{A}\frac{dA}{dt}=\frac{1}{R}\frac{dR}{dt}=H(t)=\frac{\frac{C^{''
0}_2}{2}-t}{C^{''0}_1+C^{''0}_2t-t^2}=\frac{\frac{C^{''0}_2}{2}-t}{R^2}.
\eeq
We see that, as $t\leq C^{''0}_2/2,$ $-\frac{1}{A}\frac{dA}{dt}\geq 0,$
and therefore $\frac{dS}{dt}\geq 0,$ via (6.6); hence, by virtue of
(6.14),(6.28), for the closed universe the entropy increases during the
expansion in the whole on-shell matter-dominated era, reaching its
maximum at $t=t_0,$ where $\frac{dS}{dt}=0.$
Rewriting (6.23) in the form
\beq
\left\{ \begin{array}{ccl}
R & = & \sqrt{C^{''0}_1+(C^{''0}_2)^2/4-\tilde{t}^2}, \\
A & = & \gamma /R,
\end{array} \right.
\eeq
where
\beq
\tilde{t}\equiv C^{''0}_2/2-t,
\eeq
we see that the model possesses explicit $\tilde{t}$-reversal. Since
\beq
\frac{dS}{d\tilde{t}}=-\frac{dS}{dt},
\eeq
one sees that, for $t>t_0,$ when the universe contracts, the entropy is a
growing function of $\tilde{t},$ as seen in Eqs. (6.29),(6.32).
Let us write down the formula which is valid for the closed universe and
follows from Eq. (6.6) (with $E^{'}=E$ for the on-shell matter-dominated
universe) and $A\sim 1/R:$
\beq
\frac{dS}{dt}=\frac{E}{T}H(t).
\eeq
Note also that for $A\sim 1/R,$ the energy-momentum conservation (3.5)
yields the relation (3.15) for the dust universe. Indeed, in the case of
local $O(4,1)$ invariance of the $(x^\mu ,\tau )$ manifold, it follows
from (3.11) with $A\sim 1/R$ that $R^{4\Gamma -2}\rho ={\rm const;}$ for
$\Gamma =5/4$ the latter reduces to (3.15). Similarly, in the $O(3,2)$
case, one obtains from (3.13) $R^{2\Gamma }\rho ={\rm const,}$ which
again reduces to (3.15) for $\Gamma =3/2.$
\subsection{Open dust universe}
For the open dust universe, as seen in Eqs. (6.24), as
$t\rightarrow \infty ,$
\beq
R\simeq t,\;\;\;A\simeq b-\gamma /t\rightarrow b={\rm const},
\eeq
so that this model tends asymptotically to the standard 4D cosmological
open model, for which $R\simeq t$ at large $t$ \cite{Miln}. It follows
from (6.24) that, as $t\rightarrow \infty ,$
\beq
-\frac{1}{A}\frac{dA}{dt}=-\frac{bC^{''0}_1+\gamma t}{(C^{''0}_1+t^2)^{3/
2}}\rightarrow -\frac{\gamma }{t^2}<0,
\eeq
since $A$ and $\gamma $ are of the same sign, in view of (6.17). Thus,
for the open dust universe, the entropy is a decreasing function of
cosmic time, but it can be a growing function of historical time, if $p_
\tau <0,$ similarly to the case of the flat dust universe.
\section{Concluding remarks}
We have considered 5D Kaluza-Klein type cosmological model with the fifth
coordinate being an invariant historical time $\tau .$ We have derived a
complete cosmological inflationary scenario for such a model which
distinguishes between vacuum-, off-shell matter-, and on-shell
matter-dominated eras as the solutions of the corresponding 5D
gravitational field equations. According to this scenario, the passage
from the off-shell matter-dominated era to the on-shell one occurs,
probably as a phase transition. We have studied the effect of this phase
transition on the expansion rate and found that it does not change in
either case of local $O(4,1)$- or $O(3,2)$-invariance of the extended
$(x^\mu ,\tau )$-manifold.
In contrast to the standard cosmological model in which the expansion of
the universe is adiabatic, $dS/dt=0$ \cite{fried,N}, the model considered
here does not expand adiabatically; the thermodynamic entropy is a
growing function of cosmic time for the closed universe, and can be a
growing function of historical time for the open and the flat universes.
We have obtained a complete solution of the 5D gravitational field
equations for the on-shell matter-dominated universe. We have shown that
the 5D open and closed universes tend asymptotically to the
corresponding standard 4D cosmological models, in contrast to the 5D flat
universe which does not have such a corresponding limit.
The question of the choice of the source term in the form containing $p_
5$ (like (2.3)) has received attention in the recent literature.
Mann and Vincent \cite{MV} have considered the case of local $O(4,1)$
invariance of 5D manifold and used the source term with $p_5$ involved.
The effective 4D equation of state obtained by them from the vacuum
solution of the 5D equations, $\rho =3p,\;p_5=0,$ is essentially the one
used in ref. \cite{RSS} with the quantum one-loop correction terms ($\rho
\sim -p_5\sim A^{-5})$ neglected in comparison with the classical 5D
radiation term. Wesson \cite{W} has considered the 5D version of 4D field
equations, $R_{\alpha \beta }=T_{\alpha \beta },$ with the source term
$T_{\alpha \beta }$ in the form (3.9) but only solved $R_{\alpha \beta }=
0.$ Later on, Wesson and independently Ponce de Leon \cite{WPL}
suggested that the 5D field equations may be just $R_{\alpha \beta }=0,$
and the extra terms which appears on the left-hand sides of the 5D
equations $R_{\alpha \beta }=0$ may correspond to the terms involving
matter parameters (like the density and pressure) which appear on the
right-hand sides of the 4D equations $R_{\mu \nu }=T_{\mu \nu }.$ More
recently, Ponce de Leon and Wesson \cite{PLW} have shown that a 5D theory
with no source can be cast into the form of a 4D theory with a source, in
the three-dimensionally symmetric case. In this case (when the metric is
independent of the extra coordinate, which is the case we consider in our
paper), as they have shown, the equation of state has the form of that of
radiation-like matter, $p=\rho /3.$ Hence, it is not possible, in the
case when the source term is obtained from the geometry of the higher
dimensional theory alone, to achieve the equation of state of a strongly
interacting phase (e.g., $p=\rho /4,$ which has certain experimental
evidence \cite{exp}), as we have obtained with an explicit source term,
nor to describe the inflationary epoch.
By rewriting the equation of the energy-momentum conservation, (3.5), in
the form $$\dot{\rho }+\left( 3\frac{\dot{R}}{R}+\frac{\dot{A}}{A}\right)
(\rho +p)-\frac{\dot{A}}{A}(p+p_5)=0,$$ one sees that the question of the
presence of $p_5$ in the source term for the field equations (3.2)-(3.4)
is associated with the question of whether or not the cosmological fluid
is ideal. In the case of local $O(4,1)$ invariance of $(x^\mu ,\tau )$
manifold, when $p_5=-p,$ the latter equation reduces to $$\dot{\rho }+
3\frac{\dot{R^{'}}}{R^{'}}(\rho +p)=0,$$ which is the standard equation
for ideal cosmological fluid expanding non-adiabatically due to the
scale factor $R^{'}=RA^{1/3}.$ In the other cases, $p_5=p$ $(O(3,2))$ or
$p_5=0$ \cite{MV}, the cosmological fluid is not ideal since the
energy-momentum tensor has an anisotropic pressure. Note that, as follows
from (3.5), in the case of local $O(3,2)$ invariance of $(x^\mu ,\tau )$
manifold, the cosmological fluid is ideal and expands $adiabatically$ if
$p_5=\rho ,$ implying the equality of the energy and mass densities, as
for standard radiation-like matter. Evolution of the universe filled with
such a fluid will be the subject of subsequent study.
\newpage
|
\section{Introduction}
\vspace{-0.4cm}
\indent\indent
Chiral perturbation theory ({\small $\chi$PT}) offers
a valuable guiding principle
in our attempt to relate nuclear dynamics
to the fundamental QCD.
The concept of chiral counting also gives
a clear perspective in organizing our description of
complicated nuclear dynamics.
Indeed, a new line of nuclear physics
based on {\small $\chi$PT} \,
seems to be steadily gaining ground.
In this talk,
after giving a minimal sketch of {\small $\chi$PT},
we present two examples
of the nuclear physics application of {\small $\chi$PT}.
We first discuss the latest developments
in the study of possible kaon condensation
in dense matter.
We then describe the use of {\small $\chi$PT}\,
in calculating nuclear responses
to electro-weak interaction probes.
The introduction of {\small $\chi$PT} \,follows
a generic pattern to define an effective
theory.$\!$\cite{gl84,wei90,bkm95}
Consider the vacuum-to-vacuum amplitude
in QCD in the presence of external fields
\begin{equation}
{\rm e}}\def\rmi{{\rm i}^{iZ[v,a,s,p]}=
\int[d{\mbox{\footnotesize$G$}}]
[dq] [d\bar{q}]\,
{\rm e}}\def\rmi{{\rm i}^{\rmi\int \!d^4\!x\,
{\cal L}}\def\cM{{\cal M}(q,{\bar q},G;\,v,a,s,p)}
\label{eq:ZQCDsource}
\end{equation}
where
${\cal L}}\def\cM{{\cal M}={\cal L}}\def\cM{{\cal M}^0_{\rm QCD}+\bar{q}\gamma^\mu
[v_\mu(x)-\gamma^5 a_\mu(x)]q
-\bar{q}[s(x)-ip(x)]q$.
The external fields,
$v_\mu$, $a_\mu$, $s$ and $p$,
are assigned appropriate SU(3)$\times$SU(3)
transformation properties
to make ${\cal L}}\def\cM{{\cal M}$ chiral invariant.
The effective lagrangian that describes
low-energy phenomena of QCD
($E\,\mbox{{\scriptsize \raisebox{-.9ex\Lambda_\chi\!\sim$1 GeV)
involves the Goldstone bosons and
is introduced through
\begin{equation}
{\rm e}}\def\rmi{{\rm i}^{iZ[v,a,s,p]}=
\int[d\mbox{\footnotesize{$U$}}]\,
{\rm e}}\def\rmi{{\rm i}^{\rmi\int \!d^4\!x\,
{\cal L}}\def\cM{{\cal M}_{\rm eff}(U;\,v,a,s,p)},
\label{eq:ZLeff}
\end{equation}
where
$U\equiv
\exp(i\sum_{a=1}^8\pi^a\lambda^a/f_\pi)$
with $\pi^a$ the octet pseudo-scalar mesons.
In {\small $\chi$PT}\,we expand ${\cal L}}\def\cM{{\cal M}_{\rm eff}$
in powers of $\partial_\mu/\Lambda_\chi$ and
the quark mass matrix $\cM/\Lambda_\chi$
and, for a given order of expansion,
retain all terms that are consistent with
the symmetries.
In extending this scheme
to the baryon field $N$,
we realize that $\partial_0$
acting on $N$ yields $\simm_{\mbox{\tiny N}}$,
which is not small compared with $\Lambda_\chi$.
The heavy-baryon chiral perturbation formalism
({\small HB$\chi$PF}) allows us to avoid this difficulty.$\!$\cite{jm91}
Here, instead of the ordinary Dirac field $N$
we work with $B$ defined by
$B(x)\equiv{\rm e}}\def\rmi{{\rm i}^{\rmi v\cdot x}N(x)$
with $v\sim(1,0,0,0)$,
shifting the energy reference point
from 0 to $m_{\mbox{\tiny N}}$.
If we are only concerned
with small energy-momenta $Q$
around this new origin,
the antibaryon can be ``integrated away".
${\cal L}}\def\cM{{\cal M}_{\rm eff}(B, U;\,v,a,s,p)$
describing this particle-only world
may be defined similarly to Eq. (\ref{eq:ZLeff}).
The corresponding equation of motion
for $B$ may be rewritten as coupled equations
for the large and small components $B_{\pm}$
defined by
$B_{\pm}\equiv P_{\pm}B$ with
$P_{\pm}\equiv (1\pm\!\not\!\!\mbox{\large $v$})/2$.
Eliminating $B_-$ in favor of $B_+$
leads to an equation of motion for $B_+$.
The {\small HB$\chi$PF} \,lagrangian $\cL_{\rm HB}$
is defined as an effective lagrangian
that reproduces the equation of motion
for $B_+$ and $U$.
Since $B_-\propto (Q/m_{\mbox{\tiny N}})B_+$,
$\cL_{\rm HB}$ involves expansion
in $\partial_\mu/m_{\mbox{\tiny N}}$
as well as in $\partial_\mu/\Lambda_\chi$
and $\cM/\Lambda$.
We can organize this expansion as
\begin{equation}
\cL_{\rm HB}={\cal L}}\def\cM{{\cal M}^{(1)}+{\cal L}}\def\cM{{\cal M}^{(2)}+\,\cdots\;\,;\;\;\;\;\;\;
{\cal L}}\def\cM{{\cal M}^{(\nu)}={\cal O}(Q^{\nu-1})
\end{equation}
The chiral order index $\nu$ is defined as
$\nu=d+(n/2)-2$,
where $n$ is the number of
fermion lines involved in a vertex,
and $d$ is the number of derivatives
(with $\cM\propto m_\pi^2$ counted
as two derivatives).
The explicit expression relevant
to the meson-baryon sector is\cite{bkm95}
\begin{equation}
\cL_{\rm HB}\,=\,\bar{B}_+\left[
{\cal A}}\def\cB{{\cal B}^{(1)}+{\cal A}}\def\cB{{\cal B}^{(2)}
\,+\,(\gamma_0\cB^{(1)}\gamma_0)
\frac{1}{2m_{\mbox{\tiny N}}}\cB^{(1)}\right]B_++{\cal O}(Q^2),
\label{eq:LHBAB}
\end{equation}
The leading order term is given in terms of
$u=\sqrt{U}$ and
$S_\mu=i\gamma_5\sigma_{\mu\nu}v^\nu/2$
as
\begin{eqnarray}
{\cal A}}\def\cB{{\cal B}^{(1)}&=&i(v\cdot D)+g_{\mbox{\tiny A}}(u\cdot S)\\
D_\mu&=&\partial_\mu+[u^\dagger,\partial_\mu u]/2
-i\,u^\dagger(v_\mu+a_\mu)u/2
-i\,u(v_\mu-a_\mu)u^\dagger/2
\end{eqnarray}
The expressions for higher order terms
can be found in Ref.$\!$\cite{bkm95}
Chiral counting can also be applied to
Feynmann diagrams;
the chiral order $D$ of
an irreducible Feynmann diagram
is given by\cite{wei90}
\begin{equation}
D\,=\,2-\frac{1}{2} N_E+2L-2(C-1)+\sum_i\nu_i,
\label{eq:Dcount}
\end{equation}
where $N_E$ is the number of external fermion lines,
$L$ the number of loops,
$C$ the number of disconnected parts,
and the sum runs over vertices.
\vspace{-0.2cm}
\section{Kaon Condensation in Dense Baryonic Matter}
\vspace{-0.4cm}
\indent \indent
Kaon condensation in dense baryonic matter
has been discussed
by many authors.$\!$\cite{kn86,kub93a}
According to the latest calculation,$\!$\cite{lbmr95}
the critical density $\rho_c$ for kaon condensation
is $\rho_c \approx 4\rho_0$
($\rho_0=$ normal nuclear matter density) and,
with the Brown-Rho scaling\cite{br91} included,
$\rho_c$ can be as low as $2\rho_0$.
Kaon condensation
(as we are interested in here)
is driven by the $s$-wave interactions,
unlike pion condensation
which depends on the $p$-wave interactions.
The strong $s$-wave $K$-$N$ attraction
comes partly from the so-called $\sigma$-term,
which is significantly stronger
for the kaon than for the pion.
Furthermore, the vector-meson exchange contributions
can give rise to strongly attractive s-wave interactions
for some $K$-$N$ channels,
whereas they are either repulsive
or only weakly attractive for the $\pi N$ channels.
These features motivate us
to examine the possibility of s-wave
kaon condensation.
As far as observational consequences are concerned,
a kaon condensate
(like a boson condensate in general)
could enhance significantly neutrino emission
from nascent neutron stars,
cooling them much faster.
Furthermore, the condensate
can drastically soften the equation of state
for collapsing stars.
Brown and Bethe\cite{bb94} argue that
this softening leads to proliferation
of mini blackholes, which resolves
the long-standing puzzle
that the observational value
for the ratio $R\equiv$
[\# of neutron stars]/[\# of supernova events]
is inexplicably low.
Two of the outstanding issues
facing kaon condensation are
the $m^*_N$ effect and
the off-mass-shell effects
(both to be explained below).
We wish to report here the progress we have made
on these issues over the past year.
\vspace{-0.2cm}
\subsection{The $m_{\mbox{\tiny N}}^*$ Effect}
\vspace{-0.5cm}
\indent\indent
Several authors argued
that in-medium nucleon mass reduction
could strongly hinder
kaon condensation.$\!$\cite{kub94,schetal94}
As mentioned above,
the $K$-$N$ $\sigma$-term,
$\sigma_{\mbox{\tiny KN}}\bar{\psi}\psi\bar{K}K$,
provides a significant part of the $s$-wave attraction.
The $\sigma$-term attraction
in baryonic matter
is (in the mean-field approximation)
proportional to the Lorentz scalar density
$\rho_s \equiv \,<\!\!\bar{\psi}\psi\!\!>$.
The earlier works, however,
used the approximation $\rho_s \sim \rho$,
where $\rho$ is the baryon density,
$\rho \equiv\,<\!\!\bar{\psi}\gamma_0\psi\!\!>$.
This simplifies the calculation considerably,
since $\rho$ is a conserved quantity
that can be specified as an external parameter,
whereas $\rho_s$ is known only
after the whole dynamics is solved.
For a nucleon of effective mass
$m_{\mbox{\tiny N}}^*$ and momentum ${\bf k}$,
we have
$\bar{u}_{\bf k} u_{\bf k}=
[m_{\mbox{\tiny N}}^*/(m_{\mbox{\tiny N}}^{*2}+{\bf k}^2)^{1/2}]\,
u^\dagger_{\bf k} u_{\bf k}$,
which suggests
that using $\rho$ instead of $\rho_s$
overestimates the $\sigma$-term contribution
and that this overestimation becomes
more serious for smaller values of $m_{\mbox{\tiny N}}^*$.
Detailed calculations\cite{schetal94}
based on the Walecka model\cite{sw86}
indicate that, for $m_{\mbox{\tiny N}}^*\mbox{{\scriptsize \raisebox{-.9ex 0.75\rho_0$,
the effective kaon mass $m_{\mbox{\tiny K}}^*$
does not any longer go down to zero but
levels off as $\rho$ increases,
and $m_{\mbox{\tiny K}}^*(\rho\!\rightarrow\!\infty)\mbox{{\scriptsize \raisebox{-.9ex 0.45 m_{\mbox{\tiny K}}$.
For convenience we refer to this feature
as the ``$m_{\mbox{\tiny N}}^*$ effect".
If the $m_{\mbox{\tiny N}}^*$ effect is indeed as strong
as the Walecka model suggests,
there would be no kaon condensation.
Does this argument invalidate
Lee {\it et al.}'s conclusion\cite{lbmr95}
$\rho_c=(2\!\!\sim\!\!4)\rho_0$ ?
This issue is connected
to the choice of the nucleon field.
The Walecka model uses the original Dirac field.
For systematic chiral counting, however,
it is more advantageous to work with
the heavy baryon field $B_+$,
and this is what Lee {\it et al.}\cite{lbmr95} did.
Now, for $B_+$, there is by construction
no distinction between
$\rho_s\equiv\bar{B}_+B_+$
and $\rho\equiv\bar{B}_+\gamma_0 B_+$.
In this sense Lee {\it et al.}'s approach
is free from the conventional approximation
$\rho_s\approx\rho$.
But this is of course not the whole story.
In {\small HB$\chi$PF} \,\,the effects of the $B_-$ responsible
for $\rho_s\neq\rho$
are transformed into the higher order terms in
$1/m_{\mbox{\tiny N}}$ expansion.
So we need to examine how this
$1/m_{\mbox{\tiny N}}$ expansion is handled in practice.
The lowest-order term in {\small HB$\chi$PF}\,
[{\it i.e.\ }\,${\cal A}}\def\cB{{\cal B}^{(1)}$ term in ${\cal L}}\def\cM{{\cal M}_{\rm HB}$,
Eq. (\ref{eq:LHBAB})]
applies to an infinitely heavy baryon,
and hence the $m_{\mbox{\tiny N}}^*$ effect
is totally absent here.
The next order contribution contains
$\nu=1$ terms in ordinary chiral counting
(${\cal A}}\def\cB{{\cal B}^{(2)}$ term) and terms that are first order
in $1/m_{\mbox{\tiny N}}$.
We denote the latter by ${\cal L}}\def\cM{{\cal M}_{1/m}$.
${\cal L}}\def\cM{{\cal M}_{1/m}$ consists of
the baryon kinetic energy term
${\cal L}}\def\cM{{\cal M}_{1/m}^B\equiv
\bar{B}_+(-\partial_\mu^2/2m_{\mbox{\tiny N}})B_+$
and the meson-baryon interaction part
${\cal L}}\def\cM{{\cal M}_{1/m}^{\rm int}$.
Now, to understand the calculational scheme
adopted by Lee {\it et al.},
let us rearrange ${\cal L}}\def\cM{{\cal M}_{\rm HB}$ as
\begin{eqnarray}
{\cal L}}\def\cM{{\cal M}_{\rm HB}&=&
\left(
{\cal L}}\def\cM{{\cal M}_{\rm HB}
({\rm non}\!-\!{\rm strange\;sector})
+{\cal L}}\def\cM{{\cal M}_{\rm HB}({\rm strange\;sector})
\right)
_{m_{\mbox{\tiny N}}\rightarrow\infty}
+{\cal L}}\def\cM{{\cal M}_{1/m}+\cdots
\nonumber\\
&=& \left\{
{\cal L}}\def\cM{{\cal M}_{\rm HB}({\rm non}\!-\!{\rm strange})_
{m_{\mbox{\tiny N}}\rightarrow\infty}
+{\cal L}}\def\cM{{\cal M}_{1/m}^B
+{\cal L}}\def\cM{{\cal M}_{1/m}^{\rm int}
({\rm non}\!-\!{\rm strange})
\right\}
\nonumber\\
&&\mbox{\hspace{1cm}} +
\left[
{\cal L}}\def\cM{{\cal M}_{\rm HB}({\rm strange})
_{m_{\mbox{\tiny N}}\rightarrow\infty}
+{\cal L}}\def\cM{{\cal M}_{1/m}^{\rm int}({\rm strange})
\right]
+\cdots\label{eq:Lsplit}
\end{eqnarray}
We first discuss the non-strange sector
corresponding to the terms in the curly brackets.
In the existing calculations based on {\small HB$\chi$PF}\,
the energy density for the non-strange sector
is taken from nuclear matter calculations of
the Brueckner-Hartree-Fock type.
This effectively incorporates
the $1/m_{\mbox{\tiny N}}$ correction.
In fact, since any realistic nuclear matter
calculation takes account of the change
$m_{\mbox{\tiny N}}\!\rightarrow\!m_{\mbox{\tiny N}}^*$,
the use of
the nuclear matter calculation results
allows us to go beyond
the $1/m_{\mbox{\tiny N}}$ correction.
This is in a sense a welcome feature
but there is a problem too.
In {\small HB$\chi$PF}\, the change $m_{\mbox{\tiny N}}\!\rightarrow\!m_{\mbox{\tiny N}}^*$
arises either from $(1/m_{\mbox{\tiny N}})^n$ corrections
($n\geq 2$) or from vertices with $\nu\geq 2$,
and we must deal with
a great multitude of possible terms.
By using the nuclear matter results
containing the effective mass change
one is selecting a very particular subset
of the higher order effects,
and at present there is
no clear justification for doing so.
On the other hand, the fact the change
$m_{\mbox{\tiny N}}\!\rightarrow\!m_{\mbox{\tiny N}}^*$ features importantly
in nuclear matter calculation does indicate
that one cannot simply stop at
the first correction term in $1/m_{\mbox{\tiny N}}$ expansion.
We next discuss the strangeness sector,
the terms in the square brackets
in Eq.~(\ref{eq:Lsplit}).
Here we note that ${\cal A}}\def\cB{{\cal B}^{(2)}$ terms
contained in
${\cal L}}\def\cM{{\cal M}_{\rm HB}({\rm strange})$ is of the same
chiral order ($\nu=1$)
and that the coefficients appearing in ${\cal A}}\def\cB{{\cal B}^{(2)}$
are in fact phenomenologically fixed
in such a manner that observables
for one-meson one-baryon systems be reproduced.
Then the introduction of the $1/m_{\mbox{\tiny N}}$ term
just leads to a readjustment of
these parameters.
Therefore, the $m_{\mbox{\tiny N}}^*$ effect
in the Walecka model would correspond
to terms of $\nu=2$ or higher.
Again, there are many such terms
and, for consistency, one must retain all of them.
The Walecka model represents
a particular choice of a subset,
and it remains to be seen
whether the strong $m_{\mbox{\tiny N}}^*$ effect
suggested by the model survives
a fully consistent treatment.
On the other hand,
no calculations so far done in {\small HB$\chi$PF}\,
go beyond the $1/m_{\mbox{\tiny N}}$ term
in the strangeness sector.
The only exception is
a qualitative remark by Lee {\it et al.}\cite{lbmr95}
that a multifermion term such as
$(\bar{B}_+\gamma_\mu B_+)
(\bar{B}_+B_+)\bar{K}\partial_\mu K$
can lead to a in-medium
($m_{\mbox{\tiny N}}^*$-dependent) modification
of the $K$-$N$ interaction.
This $m_{\mbox{\tiny N}}^*$ effect in fact
enhances the $K$-$N$ attraction
quite in contrast to the $m_{\mbox{\tiny N}}^*$ effect
found in the Walecka model.
Obviously, more systematic treatments
of higher order terms are required
before we can reach a solid conclusion
on the $m_{\mbox{\tiny N}}^*$ effect.
In this connection,
one may worry that
a plethora of multi-fermion vertices
that can participate in dense matter
will spoil the convergence of chiral expansion.
In fact, this does not happen as easily as
one naively expects.
According to Eq. (\ref{eq:Dcount}),
a Feynmann diagrams with a given number
of external lines $N_E$ has
a smaller value of $C$
if it contains vertices
with larger values of $n$,
thus resulting in a higher chiral order index $D$.
So, the actual contributions of vertices
with large fermion numbers
to a Feynmann diagram
are more suppressed than the chiral counting
of individual vertices would indicate.
This implies that we probably need not deal with
a tower of multi-fermion terms
to understand the $m_{\mbox{\tiny N}}^*$ effect
in the framework of {\small HB$\chi$PF}.
There have been interesting attempts
at relating the Walecka model
to {\small HB$\chi$PF}.$\!$\cite{gr94}
\vspace{-0.2cm}
\subsection{Off-Shell-Effects}
\vspace{-0.4cm}
\indent\indent
Since the main points of our discussion here
can be described more conveniently
for the pion than for the kaon,
we shall discuss the pion case.
According to the standard multiple
scattering theory,
the pion-nuclear optical potential,
or pion self-energy, is given by
\begin{equation}
\Pi=\rho\,t_{\pi A}+ \cdots,\label{eq:optical}
\end{equation}
where $t_{\pi A}$ is the $t$-matrix
describing pion scattering off
a nucleon in medium,
and the dots represent processes
involving more than a single scatterer.
The pion propagator pertaining to $t_{\pi A}$
is a full A-body nuclear hamiltonian,
not just the single nucleon hamiltonian.
Note that, in order to use $\Pi$
in the determination of
the in-medium dispersion relation for a pion,
we need information on $t_{\pi A}$
for off-shell as well as on-shell kinematics.
In the low-density limit,
we only need retain the $\rho\,t_{\pi A}$ term,
and furthermore we can replace $t_{\pi A}$
with the on-shell $t$-matrix
for free $\pi$-$N$ scattering.
Now, the issue raised
by Yabu {\it et al.}$\!$\cite{ynk93} is as follows.
Consider a toy $\pi$-$N$ lagrangian
that contains only the $\sigma$ term\footnote{
This is a highly simplified version of
the Kaplan-Nelson lagrangian.
Although recent calculations\cite{lbmr95,tw95}
take due account of energy-dependent terms of
the same chiral order as the $\sigma$ term,
our points can be explained
without those additional terms.}:
\begin{equation}
{\cal L}}\def\cM{{\cal M}_{1} = \frac{1}{2} \left[ -\phi (\Box+m_\pi^2) \phi
+ \frac{\sigma_{\pi N}}{f^2} \phi^2 {\bar{N} N} \right].\label{eQa}
\end{equation}
For ${\cal L}}\def\cM{{\cal M}_{1}$,
the $\pi$-$N$ scattering amplitude
in tree approximation is simply a constant:
$T_{\pi N}^{(1)} = \sigma_{\pi N}/f^2$.
The corresponding pion effective mass
$m_\pi^*$ (in the mean-field approximation) is
$[m_\pi^*(1)]^2 =m_\pi^2-\rho\,(\sigma_{\pi N}/f^2)$.
On the other hand,
the PCAC plus current algebra
gives the forward scattering amplitude
$T_{\pi N}^{(2)} =
[(k^2 +(k')^2-m_\pi^2)/f^2m_\pi^2]\sigma_{\pi N}$.
The corresponding $m_\pi^*$ is given by
$[m_\pi^*(2)]^2 =m_\pi^2 \,
[1 +\rho\,(\sigma_{\pi N}/m_\pi^2f^2)]\cdot
[1 +2\rho\,(\sigma_{\pi N}/m_\pi^2f^2)]^{-1}$.
Although $m_\pi^*(1)$ and $m_\pi^*(2)$
are identical for low densities,
they behave very differently for large values of $\rho$.
In particular, $m_\pi^*(2) \!\rightarrow \!m_\pi/\sqrt{2}$ as
$\rho \!\!\rightarrow \!\!\infty$.
Yabu {\it et al.}, who pointed out this discrepancy,
argued that the existing calculational frameworks
did not allow one to resolve this problem.
It behooves to remember here
the following general points:
(i) The formal definition of
$m_\pi^*$ is a value of the energy variable
$\omega$ for which
the exact in-medium Green's function
$G_\rho(x;\phi)=
<\!\rho|T\phi(x)\phi(0)|\rho\!>$
develops a pole (for zero momentum);
(ii) For a given lagrangian,
the physical observable $m_\pi^*$
should not depend on the definition
of interpolating fields $\phi$;
(iii) Although off-mass-shell $\pi$-$N$ amplitudes
vary for different choices of $\phi$,
this variation should not affect any observables
including $m_\pi^*$;
(iv) Although off-shell $\pi$-$N$ amplitudes are
unphysical in the sense of (iii) and also in that
they cannot be observed in $\pi$-$N$ scattering,
they do constitute ingredients of
larger Feynmann diagrams;
(v) The statements (i)$\sim$(iii) hold true
only if the whole calculation is done exactly.
This last point is trivial but
nonetheless worth emphasizing.
Now, within the framework of
the leading order optical potential,
the variance
between $m_\pi^*(1)$ and $m_\pi^*(2)$
is a direct consequence of the
fact that $T_{\pi N}^{(1)}$ and $T_{\pi N}^{(2)}$
have different off-shell behaviors.
Referring to the above general statements,
one could ask
whether this is a manifestation of different dynamics,
or just a spurious off-shell effect
that fails to disappear because of
the approximation used.
Yabu {\it et al.}\,favored the first possibility,
conjecturing that different treatments of
multi-fermion terms are responsible
for the different behaviors of
$T_{\pi N}^{(1)}$ and $T_{\pi N}^{(2)}$.
This interpretation, however, was criticized
by Lee {\it et al.}$\!$\cite{lbmr95}
and by Thorsson and
Wirzba (TW).$\!$\cite{tw95}
TW show explicitly that,
starting from the same ${\cal L}}\def\cM{{\cal M}_{HB}$,
one can derive either of
$T_{\pi N}^{(1)}$ and $T_{\pi N}^{(2)}$
by adding to ${\cal L}}\def\cM{{\cal M}_{\rm HB}$
different pseudoscalar source terms.
This ensures that,
provided one can calculate
$G_\rho(x;\phi)=
<\!\rho|T\phi(x)\phi(0)|\rho\!>$
{\underline{exactly}},
one would get the same $m_\pi^*$
regardless of whether one uses
$T_{\pi N}^{(1)}$ or $T_{\pi N}^{(2)}$.
Beautiful !!
(Please note, however, the underline
attached to ``exactly".)
In practice, we must adopt some approximation,
the crudest and most commonly used approximation
being $\Pi\approx \rho\,t_{\pi\,N}$.
In these approximate calculations,
choice between $T_{\pi N}^{(1)}$ and
$T_{\pi N}^{(2)}$ does matter,
and TW's formal proof is not of immediate help
in making this choice.
We must mention here, however,
another important point made by TW.
TW demonstrates that,
within the mean-field approximation,
the use of the effective action leads
to the identical dispersion relation for
an in-medium pion regardless of
different choices of the pseudoscalar source.
This is a remarkable result,
but it seems important to
examine to what extent this theorem
is tied to the mean field approximation.
In fact, if TW's result is valid
beyond the mean filed approximation,
that would give a tremendous impact
to the ``standard" multiple scattering formalism.
We would be forced to conclude that
the obvious off-shell dependence
exhibited by the leading term
in the Watson expansion is spurious
(at least for a system the dynamics of which
is strongly constrained by chiral symmetry).
This point deserves a serious investigation
quite apart from the specific problem of
meson condensation.
\section{Nuclear Responses to Electro-Weak Probes}
\vspace{-0.4cm}
\indent \indent
The nuclear hamiltonian is normally taken to be
$H_N = \sum_{i=1}^{A} T_i + \sum_{i,j}^{A} V_{ij}$,
where $T_i$ is the nucleon kinetic energy,
and $V_{ij}$ is the ``realistic" $N$-$N$ potential.
Arriving at $H_N$ starting
from the fundamental QCD description involves:
(i) translating the quark and gluon degrees of freedom
into the effective degrees of freedom of hadrons;
(ii) truncating the Hilbert space of hadrons
down to that of non-relativistic nucleons
interacting via potentials.
The {\small $\chi$PT} \,allows us to carry out (i) and (ii)
in a well-defined way,
preserving the basic chiral properties of QCD.
Construction of the realistic $N$-$N$ potentials
based on {\small $\chi$PT} \,was described
by Weinberg\cite{wei90}
and by van Kolck {\it et al.}\cite{kol92}
These {\small $\chi$PT}\, potentials can reproduce
the $N$-$N$ observables almost as satisfactorily
as the conventional boson-exchange potentials
which contain many {\it ad hoc} parameters.
In the truncated nucleonic space,
nuclear responses to external probes
such as electromagnetic and weak currents
involve not only single-nucleonic terms
(= impulse approximation terms)
but also multi-nucleonic contributions
named the exchange currents.
Here again, {\small $\chi$PT} \, provides a systematic framework
for organizing exchange-current contributions
according to their chiral
counting orders.$\!$\cite{rho91,pmr93,pmr95}
A problem in testing the exchange currents
in complex nuclei
is that exact solutions
for the $A$-body Schr\"{o}dinger equation
$H_N\Psi= E\Psi$ are hard to obtain
and therefore we are forced to work
with truncated model wave functions $\Psi_0$.
If the matrix element of
a nuclear operator ${\cal O}$ is calculated
using model wave functions,
then
$\mbox{$<\!\Psi^f|{\cal O}|\Psi^i\!>$} \neq
<\!\Psi_0^f|{\cal O}|\Psi_0^i\!>$.
This deviation represents the core-polarization effect.
The core polarization effects
need to be carefully sorted out
before one can identify
the exchange currents effects.
Despite this non-trivial aspect,
there is growing evidence
that supports the {\small $\chi$PT} \,derivation of exchange currents.
The best example is
the nuclear axial-charge operator $A_0$.
Warburton {\it et al.}'s
systematic analyses\cite{war91,wt94}
of the first-forbidden $\beta$ transitions
indicate that the ratio of
the exchange-current contribution
to the 1-body contribution is
$\delta_{\rm{mec}}\equiv
\langle A^0({\rm mec})\rangle/
\langle\!A^0(1{\mbox{-}}{\rm body})\rangle
\!=\!0.6 \sim 0.8$.
(The semi-empirical method used
in these analyses largely eliminates ambiguities
due to the core-polariation effects.)
The leading-order {\small $\chi$PT} \,term,
{\it i.e.\ } the soft-pion exchange
term,$\!$\cite{kdr78,rho91}
can explain the bulk of $\delta_{\rm{mec}}$,
and the next-order
{\small $\chi$PT}\,term\cite{pmr93,ptk94}
gives an additional $\sim$10\% enhancement,
bringing the theoretical value close to
the empirical value.
It is informative to compare
the above results with those obtained
in the conventional
meson-exchange approach.$\!$\cite{krt92,tow92}
Using the ``hard-pion formalism"
in conjunction with the lagrangian
that engenders the phenemenological
$N$-$N$ interactions,
Towner\cite{tow92} finds
that the pion-exchange contribution
is reduced significantly
by the phenomenological form factors,
but the reduction is largely compensated
by heavy-meson pair graphs.
The net result is:
$\delta_{\rm{mec}}^{\rm Towner}
\sim\delta_{\rm{mec}}^{\rm CPhT}$.
In fact, the former is slightly larger,
but this small difference is qualitatively
understood as follows.
The largest heavy-meson pair contributions
come from $\sigma$ and $\omega$ mesons,
and the $\sigma$-meson contribution
can be effectively rewritten
as the $1$-body term with the nucleon mass
replaced by an effective mass.$\!$\cite{dt87}
Thus the phenomenological $\sigma$-meson
plays a role similar to the BR scaling.$\!$\cite{br91}
Meanwhile, in {\small $\chi$PT},
the BR scaling is attributable
to multi-fermion terms
which have higher chiral orders than
those appearing in the next-to-leading-order
calculation of Park {\it et al.}$\!$\cite{pmr93,ptk94}
Then, we should qualitatively expect
$\delta_{\rm{mec}}$
obtained by Park {\it et al.} \,
to be somewhat smaller than
$\delta_{\rm{mec}}^{\rm Towner}$.
The above example demonstrates
the usefulness of CPT\,
in organizing complicated exchange-current
contributions in a systematic manner.
For the two-nucleon systems
we can obtain exact solutions
for $H_N\Psi= E\Psi$,
avoiding thereby the core-polarization problem.
The A=2 systems therefore
provide a clean case for checking the validity of
the standard calculational framework
based on the nucleonic Schr\"{o}dinger equation
supplemented with the exchange currents.
A beautiful test is found
in radiative capture of a thermal neutron
by a proton:
$n+p \rightarrow d+\gamma$.
The observed capture rate for this process is
$\sigma_{\rm exp}=334.2 \pm 0.5$mb,
which is $\sim$10\% larger than the IA prediction
$\sigma_{IA}=302.5\pm4.0$mb.
According to Riska and Brown,$\!$\cite{rb72}
the one-pion exchange current derived
from the low-energy theorem
can account for $\sim$70\% of the
missing capture rate.
Recognizing that this contribution represents
the leading order term in {\small $\chi$PT},
it is of great interest to examine
what the next-order term will do.
Park {\it et al.}'s recent calculation\cite{pmr95}
that includes
the next-to-leading order terms gives
$\sigma=334\pm2$mb,
in perfect agreement with experiment.
(Another impressive success of
the exchange current calculations
based on the low-energy theorem
is known for the
$e+d\rightarrow e+p+n$ reaction,
see {\it e.g.} \,Ref.\cite{fm89}.)
Our last topic is neutrino reactions on
the deuteron.
The recent developments
in the solar neutrino problem
have further enhanced the importance of
the MSW effect
as a possible mechanism to explain
the observed energy dependence
of the solar neutrino deficit.$\!$\cite{sch95}
The SNO heavy-water
\v{C}erenkov counter\cite{ardetal87}
can provide crucial information on this issue
because of its capability
to register the charged-
and neutral-current reactions
simultaneously but separately.
The SNO is also expected to be highly useful
for studying supernova neutrinos.
The neutrino-deuteron reactions
relevant to the SNO are:
$\nu + d \rightarrow \nu' + n + p$,
$\bar{\nu} + d \rightarrow \bar{\nu} + n + p$,
$\nu_e + d \rightarrow e^- + p + p$ and
$\bar \nu_e + d \rightarrow e^+ + n + n$.
Obviously, one needs reliable estimates
of the cross sections
for these reactions
to extract useful astrophysical information
from SNO data.
The above discussion indicates that
one can have enough confidence
in the calculational framework
that uses the nucleonic Schr\"{o}dinger equation
with realistic $N$-$N$ interactions
supplemented with the exchange currents.
Although one may eventually be able to
obtain all the ingredients from {\small $\chi$PT},
it is reasonable to use phenomenological input.
There have been several calculations
of this type\cite{tkk90,kk92},
and the best available estimates (in our opinion)
have been given by Kohyama {\it et al.}$\!$\cite{kk92}
\vspace{0.2cm}
|
\section{Introduction}
\label{intro}
We will find a formula for the number $I(n)$ of intersection points
formed inside a regular $n$-gon by its diagonals.
The case $n=30$ is depicted in Figure~\ref{30gon}.
For a {\em generic} convex $n$-gon, the answer would be $n \choose 4$,
because every four vertices would be the endpoints of a unique pair
of intersecting diagonals.
But $I(n)$ can be less, because in a regular $n$-gon
it may happen that three or more diagonals meet at an interior point,
and then some of the $n \choose 4$ intersection points will coincide.
In fact, if $n$ is even and at least~6,
$I(n)$ will always be less than $n \choose 4$,
because there will be $n/2 \ge 3$ diagonals meeting at the center point.
It will result from our analysis that for $n>4$,
the maximum number of diagonals of the regular $n$-gon
that meet at a point other than the center is
$$\begin{array}{rl}
2 & \text{if $n$ is odd}, \\
3 & \text{if $n$ is even but not divisible by 6}, \\
5 & \text{if $n$ is divisible by 6 but not 30, and}, \\
7 & \text{if $n$ is divisible by 30}.
\end{array}$$
with two exceptions: this number is~2 if $n=6$, and~4 if $n=12$.
In particular, it is impossible to have~8 or more diagonals
of a regular $n$-gon meeting at a point other than the center.
Also, by our earlier remarks, the fact that no three diagonals
meet when $n$ is odd will imply that $I(n)={n \choose 4}$ for odd $n$.
\begin{figure}[p]
\centerline{\psfig{file=fig1.ps,width=6.5in}}
\caption{ The 30-gon with its diagonals.
There are 16801 interior intersection
points: 13800 two line intersections, 2250 three line intersections,
420 four line intersections, 180 five line intersections, 120 six
line intersections, 30 seven line intersections, and 1 fifteen
line intersection.}
\label{30gon}
\end{figure}
A careful analysis of the possible configurations of three diagonals
meeting will provide enough information to permit us in theory
to deduce a formula for $I(n)$.
But because the explicit description of these configurations is so complex,
our strategy will be instead to use this information
to deduce only the {\em form} of the answer,
and then to compute the answer for enough small $n$
that we can determine the result precisely.
The computations are done in Mathematica, Maple and C,
and annotated source codes can be obtained via anonymous ftp
at {\verb+http://math.berkeley.edu/~poonen+}.
In order to write the answer in a reasonable form, we define
$$\delta_m(n) = \begin{cases}
1 & \text{if $n \equiv 0 \pmod m$,} \\
0 & \text{otherwise.}
\end{cases}$$
\begin{theorem}
\label{countintersections}
For $n \ge 3$,
\begin{eqnarray*}
I(n) & = & {n \choose 4}
+ (-5 n^3 + 45 n^2 - 70 n + 24)/24 \cdot \delta_2(n)
- (3n/2) \cdot \delta_4(n) \\
& & \mbox{} + (-45 n^2 + 262n)/6 \cdot \delta_6(n)
+ 42n \cdot \delta_{12}(n)
+ 60 n \cdot \delta_{18}(n) \\
& & \mbox{} + 35n \cdot \delta_{24}(n)
- 38n \cdot \delta_{30}(n)
- 82n \cdot \delta_{42}(n)
- 330n \cdot \delta_{60}(n) \\
& & \mbox{} - 144n \cdot \delta_{84}(n)
- 96n \cdot \delta_{90}(n)
- 144n \cdot \delta_{120}(n)
- 96n \cdot \delta_{210}(n).
\end{eqnarray*}
\end{theorem}
Further analysis, involving Euler's formula $V-E+F=2$,
will yield a formula for the number $R(n)$ of regions
that the diagonals cut the $n$-gon into.
\begin{theorem}
\label{countregions}
For $n \ge 3$,
\begin{eqnarray*}
R(n) & = & (n^4 - 6 n^3 + 23 n^2 - 42 n + 24)/24 \\
& & \mbox{}
+ (-5 n^3 + 42 n^2 - 40 n - 48)/48 \cdot \delta_2(n)
- (3n/4) \cdot \delta_4(n) \\
& & \mbox{} + (-53 n^2 + 310n)/12 \cdot \delta_6(n)
+ (49n/2) \cdot \delta_{12}(n)
+ 32 n \cdot \delta_{18}(n) \\
& & \mbox{} + 19n \cdot \delta_{24}(n)
- 36n \cdot \delta_{30}(n)
- 50n \cdot \delta_{42}(n)
- 190n \cdot \delta_{60}(n) \\
& & \mbox{} - 78n \cdot \delta_{84}(n)
- 48n \cdot \delta_{90}(n)
- 78n \cdot \delta_{120}(n)
- 48n \cdot \delta_{210}(n).
\end{eqnarray*}
\end{theorem}
These problems have been studied by many authors before,
but this is apparently the first time the
correct formulas have been obtained.
The Dutch mathematician Gerrit Bol~\cite{bol}
gave a complete solution in~1936,
except that a few of the coefficients in his formulas are wrong.
(A few misprints and omissions in Bol's paper
are mentioned in~\cite{rigby2}.)
The approaches used by us and Bol are similar in many ways.
One difference (which is not too substantial)
is that we work as much as possible with roots of unity
whereas Bol tended to use more trigonometry
(integer relations between sines of rational multiples of $\pi$).
Also, we relegate much of the work to the computer,
whereas Bol had to enumerate the many cases by hand.
The task is so formidable that it is amazing to us
that Bol was able to complete it,
and at the same time not so surprising that it would contain a few errors!
Bol's work was largely forgotten.
In fact, even we were not aware of his paper
until after deriving the formulas ourselves.
Many other authors in the interim solved special cases of the problem.
Steinhaus~\cite{steinhausproblem} posed the problem
of showing that no three diagonals meet internally when $n$ is prime,
and this was solved by Croft and Fowler~\cite{croft}.
(Steinhaus also mentions this in~\cite{steinhaus},
which includes a picture of the 23-gon and its diagonals.)
In the 1960s, Heineken~\cite{heineken} gave a delightful argument
which generalized this to all odd $n$,
and later he~\cite{heineken2} and Harborth~\cite{harborth}
independently enumerated all three-diagonal intersections
for $n$ not divisible by 6.
The classification of three-diagonal intersections
also solves Colin Tripp's problem~\cite{tripp}
of enumerating ``adventitious quadrilaterals,''
those convex quadrilaterals for which the angles
formed by sides and diagonals are all rational multiples of $\pi$.
See Rigby's paper~\cite{rigby2} or the summary~\cite{rigby} for details.
Rigby, who was aware of Bol's work, mentions that Monsky and Pleasants
also each independently classified all three-diagonal intersections
of regular $n$-gons.
Rigby's papers partially solve Tripp's further problem of proving
the existence of all adventitious quadrangles using only
elementary geometry; i.e., without resorting to trigonometry.
All the questions so far have been in the Euclidean plane.
What happens if we count the interior intersections made by the diagonals
of a hyperbolic regular $n$-gon?
The answers are exactly the same, as pointed out in~\cite{rigby2},
because if we use Beltrami's representation of points of the
hyperbolic plane by points inside a circle in the Euclidean plane,
we can assume that the center of the hyperbolic $n$-gon corresponds
to the center of the circle, and then the hyperbolic $n$-gon with its
diagonals looks in the model exactly like a Euclidean regular $n$-gon
with its diagonals.
It is equally easy to see that the answers will be the same in
elliptic geometry.
\section{When do three diagonals meet?}
\label{threemeet}
We now begin our derivations of the formulas for $I(n)$ and $R(n)$.
The first step will be to find a criterion for the concurrency
of three diagonals.
Let $A,B,C,D,E,F$ be six distinct points in order on
a unit circle dividing up the circumference into arc lengths
$u,x,v,y,w,z$ and assume that the three chords $AD, BE, CF$
meet at $P$ (see Figure~\ref{geometry}).
\begin{figure}
\centerline{\psfig{file=fig2.ps,width=3in}}
\caption{\vrule depth 20pt width0pt}
\label{geometry}
\end{figure}
By similar triangles, $AF/CD = PF/PD$,
$BC/EF = PB/PF$,
$DE/AB = PD/PB$.
Multiplying these together yields
\begin{displaymath}
(AF \cdot BC \cdot DE)
/(CD \cdot EF \cdot AB)
= 1,
\end{displaymath}
and so
\begin{equation}
\label{product}
\sin(u/2) \sin(v/2) \sin(w/2)
= \sin(x/2) \sin(y/2) \sin(z/2).
\end{equation}
Conversely, suppose six distinct points $A,B,C,D,E,F$
partition the circumference of a unit circle into
arc lengths $u,x,v,y,w,z$ and suppose that~(\ref{product})
holds. Then the three diagonals
$AD, BE, CF$ meet in a single point which we see as follows.
Let lines $AD$ and $BE$ intersect at $P_0$ .
Form the line through $F$ and $P_0$
and let $C'$ be the other intersection point of $FP_0$
with the circle. This partitions the circumference into
arc lengths $u,x,v',y',w,z$. As shown above, we have
\begin{displaymath}
\sin(u/2) \sin(v'/2) \sin(w/2)
= \sin(x/2) \sin(y'/2) \sin(z/2)
\end{displaymath}
and since we are assuming that~(\ref{product}) holds
for $u,x,v,y,w,z$ we get
\begin{displaymath}
\frac{\sin(v'/2)}{\sin(y'/2)} = \frac{\sin(v/2)}{\sin(y/2)}.
\end{displaymath}
Let $\alpha=v + y = v' + y'$. Substituting $v=\alpha-y$,
$v'=\alpha-y'$ above we get
\begin{displaymath}
\frac{\sin(\alpha/2) \cos(y'/2)
- \cos(\alpha/2) \sin(y'/2)}{\sin(y'/2)}
= \frac{\sin(\alpha/2) \cos(y/2)
- \cos(\alpha/2) \sin(y/2)}{\sin(y/2)}
\end{displaymath}
and so
\begin{displaymath}
\cot(y'/2) = \cot(y/2).
\end{displaymath}
Now $0<\alpha/2<\pi$, so $y=y'$ and hence $C=C'$. Thus,
the three diagonals $AD, BE, CF$ meet at a single point.
So~(\ref{product}) gives a necessary and sufficient
condition (in terms of arc lengths) for the chords $AD, BE, CF$
formed by six distinct points $A,B,C,D,E,F$
on a unit circle to meet at a single point.
In other words, to give an explicit answer to the question
in the section title, we need to characterize the positive
rational solutions to
\begin{eqnarray}
\label{product2}
\sin(\pi U) \sin(\pi V) \sin(\pi W)
& = & \sin(\pi X) \sin(\pi Y) \sin(\pi Z) \\
\nonumber U+V+W+X+Y+Z & = & 1.
\end{eqnarray}
(Here $U=u/(2\pi)$, etc.)
This is a trigonometric diophantine equation in the sense of~\cite{conway},
where it is shown that in theory, there is a finite computation
which reduces the solution of such equations to
ordinary diophantine equations.
The solutions to the analogous equation with only two sines
on each side are listed in~\cite{myerson}.
If in~(\ref{product2}), we substitute
$\sin(\theta)=(e^{i \theta}-e^{-i \theta})/(2i)$,
multiply both sides by $(2i)^3$, and expand,
we get a sum of eight terms on the left equalling
a similar sum on the right, but two terms on the left
cancel with two terms on the right since $U+V+W=1-(X+Y+Z)$, leaving
\begin{eqnarray*}
\lefteqn{-e^{i\pi (V+W-U)} + e^{-i\pi(V+W-U)}
- e^{i\pi(W+U-V)} + e^{-i\pi(W+U-V)}
- e^{i\pi(U+V-W)} + e^{-i\pi(U+V-W)} = } \\
&& -e^{i\pi (Y+Z-X)} + e^{-i\pi(Y+Z-X)} - e^{i\pi(Z+X-Y)}
+ e^{-i\pi(Z+X-Y)} - e^{i\pi(X+Y-Z)} + e^{-i\pi(X+Y-Z)}.
\end{eqnarray*}
If we move all terms to the left hand side,
convert minus signs into $e^{-i \pi}$, multiply by $i=e^{i\pi/2}$, and let
\begin{eqnarray*}
\alpha_1 & = & V+W-U-1/2 \\
\alpha_2 & = & W+U-V-1/2 \\
\alpha_3 & = & U+V-W-1/2 \\
\alpha_4 & = & Y+Z-X+1/2 \\
\alpha_5 & = & Z+X-Y+1/2 \\
\alpha_6 & = & X+Y-Z+1/2,
\end{eqnarray*}
we obtain
\begin{equation}
\label{thetwelve}
\sum_{j=1}^6 e^{i \pi \alpha_j}
+ \sum_{j=1}^6 e^{-i \pi \alpha_j} = 0,
\end{equation}
in which $\sum_{j=1}^6 \alpha_j = U+V+W+X+Y+Z = 1$.
Conversely, given rational numbers
$\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6$
(not necessarily positive) which sum to 1 and satisfy~(\ref{thetwelve}),
we can recover $U,V,W,X,Y,Z$, (for example, $U=(\alpha_2+\alpha_3)/2+1/2$),
but we must check that they turn out positive.
\section{Zero as a sum of 12 roots of unity}
\label{twelve}
In order to enumerate the solutions to~(\ref{product2}), we are led,
as in the end of the last section, to classify the ways
in which 12 roots of unity can sum to zero.
More generally, we will study relations of the form
\begin{equation}
\label{relation}
\sum_{i=1}^k a_i \eta_i = 0,
\end{equation}
where the $a_i$ are positive integers, and the $\eta_i$
are distinct roots of unity.
(These have been studied previously by Schoenberg~\cite{schoenberg},
Mann~\cite{mann}, Conway and Jones~\cite{conway}, and others.)
We call $w(S)=\sum_{i=1}^k a_i$ the {\em weight} of the relation $S$.
(So we shall be particularly interested in relations of weight~12.)
We shall say the relation~(\ref{relation}) is {\em minimal}
if it has no nontrivial subrelation; i.e., if
$$\sum_{i=1}^k b_i \eta_i = 0, \quad a_i \ge b_i \ge 0$$
implies either $b_i=a_i$ for all $i$ or $b_i=0$ for all $i$.
By induction on the weight, any relation can be represented
as a sum of minimal relations (but the representation need not be unique).
Let us give some examples of minimal relations.
For each $n \ge 1$, let $\zeta_n=\exp(2 \pi i/n)$
be the standard primitive $n$-th root of unity.
For each prime $p$, let $R_p$ be the relation
$$1 + \zeta_p + \zeta_p^2 + \cdots + \zeta_p^{p-1} = 0.$$
Its minimality follows from the irreducibility of the cyclotomic polynomial.
Also we can ``rotate'' any relation by multiplying
through by an arbitrary root of unity to obtain a new relation.
In fact, Schoenberg~\cite{schoenberg} proved that every relation
(even those with possibly negative coefficients)
can be obtained as a linear combination with positive
and negative integral coefficients of the $R_p$ and their rotations.
But we are only allowing positive combinations,
so it is not clear that these are enough to generate all relations.
In fact it is not even true!
In other words, there are other minimal relations.
If we subtract $R_3$ from $R_5$, cancel the 1's and
incorporate the minus signs into the roots of unity,
we obtain a new relation
\begin{equation}
\label{r5r3}
\zeta_6 + \zeta_6^{-1} + \zeta_5
+ \zeta_5^2 + \zeta_5^3 + \zeta_5^4 = 0,
\end{equation}
which we will denote $(R_5:R_3)$.
In general, if $S$ and $T_1,T_2,\ldots,T_j$ are relations,
we will use the notation $(S:T_1,T_2,\ldots,T_j)$
to denote any relation obtained by rotating the $T_i$
so that each shares exactly one root of unity with $S$
which is different for each $i$, subtracting them from $S$,
and incorporating the minus signs into the roots of unity.
For notational convenience, we will write $(R_5:4R_3)$
for $(R_5:R_3,R_3,R_3,R_3)$, for example.
Note that although $(R_5:R_3)$ denotes unambiguously (up to rotation)
the relation listed in~(\ref{r5r3}), in general there will be
many relations of type $(S:T_1,T_2,\ldots,T_j)$ up to rotational equivalence.
Let us also remark that including $R_2$'s in the list of $T$'s has no effect.
It turns out that recursive use of the construction above
is enough to generate all minimal relations of weight up to 12.
These are listed in Table~\ref{minimals}.
The completeness and correctness of the table will be
proved in Theorem~\ref{table} below.
Although there are 107 minimal relations up to rotational equivalence,
often the minimal relations within one of our classes are Galois conjugates.
For example, the two minimal relations of type $(R_5:2R_3)$ are
conjugate under $\operatorname{Gal}({\Bbb Q}(\zeta_{15})/{\Bbb Q})$, as pointed out in~\cite{mann}.
The minimal relations with $k \le 7$ ($k$ defined as in~(\ref{relation}))
had been previously catalogued in~\cite{mann},
and those with $k \le 9$ in~\cite{conway}.
In fact, the $a_i$ in these never exceed~1,
so these also have weight less than or equal to~9.
\begin{table}
\begin{center}
\begin{tabular}{c|c|c|}
Weight & Relation type &
Number of relations of that type \\ \hline \hline
2 & $R_2$ & 1 \\ \hline
3 & $R_3$ & 1 \\ \hline
5 & $R_5$ & 1 \\ \hline
6 & $(R_5:R_3)$ & 1 \\ \hline
7 & $(R_5:2R_3)$ & 2 \\ \cline{2-3}
& $R_7$ & 1 \\ \hline
8 & $(R_5:3R_3)$ & 2 \\ \cline{2-3}
& $(R_7:R_3)$ & 1 \\ \hline
9 & $(R_5:4R_3)$ & 1 \\ \cline{2-3}
& $(R_7:2R_3)$ & 3 \\ \hline
10 & $(R_7:3R_3)$ & 5 \\ \cline{2-3}
& $(R_7:R_5)$ & 1 \\ \hline
11 & $(R_7:4R_3)$ & 5 \\ \cline{2-3}
& $(R_7:R_5,R_3)$ & 6 \\ \cline{2-3}
& $(R_7:(R_5:R_3))$ & 6 \\ \cline{2-3}
& $R_{11}$ & 1 \\ \hline
12 & $(R_7:5R_3)$ & 3 \\ \cline{2-3}
& $(R_7:R_5,2R_3)$ & 15 \\ \cline{2-3}
& $(R_7:(R_5:R_3),R_3)$ & 36 \\ \cline{2-3}
& $(R_7:(R_5:2R_3))$ & 14 \\ \cline{2-3}
& $(R_{11}:R_3)$ & 1 \\ \hline
\end{tabular}
\end{center}
\caption{The 107 minimal relations of weight up to 12.}
\label{minimals}
\end{table}
\begin{theorem}
\label{table}
Table~\ref{minimals} is a complete listing of the minimal
relations of weight up to 12 (up to rotation).
\end{theorem}
The following three lemmas will be needed in the proof.
\begin{lemma}
\label{mann1}
If the relation~(\ref{relation}) is minimal,
then there are distinct primes $p_1<p_2<\cdots<p_s \le k$
so that each $\eta_i$ is a $p_1 p_2 \cdots p_s$-th root of unity,
after the relation has been suitably rotated.
\end{lemma}
\begin{pf}
This is a corollary of Theorem~1 in~\cite{mann}.
\end{pf}
\begin{lemma}
\label{2p}
The only minimal relations (up to rotation) involving only
the $2p$-th roots of unity, for $p$ prime, are $R_2$ and $R_p$.
\end{lemma}
\begin{pf}
Any $2p$-th root of unity is of the form $\pm \zeta^i$.
If both $+\zeta^i$ and $-\zeta^i$ occurred in the same relation,
then $R_2$ occurs as a subrelation.
So the relation has the form
$$\sum_{i=0}^{p-1} c_i \zeta_p^i = 0$$
By the irreducibility of the cyclotomic polynomial,
$\{1,\zeta_p,\ldots,\zeta_p^{p-1}\}$ are independent over ${\Bbb Q}$
save for the relation that their sum is zero, so all the $c_i$ must be equal.
If they are all positive, then $R_p$ occurs as a subrelation.
If they are all negative, then $R_p$ rotated by -1 (i.e., 180 degrees)
occurs as a subrelation.
\end{pf}
\begin{lemma}
\label{pigeonhole}
Suppose $S$ is a minimal relation, and $p_1<p_2<\cdots<p_s$
are picked as in Lemma~\ref{mann1} with $p_1=2$ and $p_s$ minimal.
If $w(S)<2 p_s$, then $S$ (or a rotation) is of the
form $(R_{p_s}:T_1,T_2,\ldots,T_j)$ where the $T_i$ are minimal relations
not equal to $R_2$ and involving only $p_1 p_2 \cdots p_{s-1}$-th
roots of unity, such that $j<p_s$ and
$$\sum_{i=1}^j [w(T_i)-2] = w(S) - p_s.$$
\end{lemma}
\begin{pf}
Since every $p_1 p_2 \ldots p_s$-th root of unity is uniquely expressible
as the product of a $p_1 p_2 \ldots p_{s-1}$-th root of unity
and a $p_s$-th root of unity, the relation can be rewritten as
\begin{equation}
\label{peel}
\sum_{i=0}^{p_s-1} f_i \zeta_{p_s}^i = 0,
\end{equation}
where each $f_i$ is a sum of $p_1 p_2 \ldots p_{s-1}$-th roots of unity,
which we will think of as a sum (not just its value).
Let $K_m$ be the field obtained by adjoining the $p_1 p_2 \ldots p_m$-th
roots of unity to ${\Bbb Q}$.
Since
$[K_s:K_{s-1}]=\phi(p_1 p_2 \cdots p_s)/\phi(p_1 p_2 \cdots p_{s-1})
=\phi(p_s)=p_s-1$,
the only linear relation satisfied
by $1,\zeta_{p_s},\ldots,\zeta_{p_s}^{p_s-1}$ over $K_{s-1}$
is that their sum is zero.
Hence~(\ref{peel}) forces the values of the $f_i$ to be equal.
The total number of roots of unity in all the $f_i$'s is $w(S)<2 p_s$,
so by the pigeonhole principle, some $f_i$ is zero or
consists of a single root of unity.
In the former case, each $f_j$ sums to zero,
but at least two of these sums contain at least one root of unity,
since otherwise $s$ was not minimal,
so one of these sums gives a subrelation of $S$, contradicting its minimality.
So some $f_i$ consists of a single root of unity.
By rotation, we may assume $f_0=1$.
Then each $f_i$ sums to 1, and if it is not simply the single root
of unity~1, the negatives of the roots of unity in $f_i$ together
with~1 form a relation $T_i$ which is not $R_2$ and involves
only $p_1 p_2 \cdots p_{s-1}$-th roots of unity, and it is clear
that $S$ is of type $(R_{p_s}:T_{i_1},T_{i_2},\ldots,T_{i_j})$.
If one of the $T$'s were not minimal, then it could be decomposed
into two nontrivial subrelations, one of which would not share a root
of unity with the $R_{p_s}$, and this would give a nontrivial subrelation
of $S$, contradicting the minimality of $S$.
Finally, $w(S)$ must equal the sum of the weights of $R_{p_s}$
and the $T$'s, minus $2j$ to account for the roots of unity that
are cancelled in the construction
of $(R_{p_s}:T_{i_1},T_{i_2},\ldots,T_{i_j})$.
\end{pf}
\begin{pf*}{Proof of Theorem~\ref{table}}
We will content ourselves with proving that every relation of
weight up to 12 can be decomposed into a sum of the ones listed
in Table~\ref{minimals}, it then being straightforward to check that
the entries in the table are distinct, and that none of them can be
further decomposed into relations higher up in the table.
Let $S$ be a minimal relation with $w(S) \le 12$.
Pick $p_1<p_2<\cdots<p_s$ as in Lemma~\ref{mann1} with $p_1=2$
and $p_s$ minimal.
In particular, $p_s \le 12$, so $p_s=2$,3,5,7, or $11$.
\medskip
\noindent{\em Case 1: } $p_s \le 3$
Here the only minimal relations are $R_2$ and $R_3$, by Lemma~\ref{2p}.
\medskip
\noindent{\em Case 2: } $p_s=5$
If $w(S)<10$, then we may apply Lemma~\ref{pigeonhole} to
deduce that $S$ is of type
$(R_5:T_1,T_2,\ldots,T_j)$
Each $T$ must be $R_3$ (since $p_{s-1} \le 3$), and $j=w(S)-5$ by
the last equation in Lemma~\ref{pigeonhole}.
The number of relations of type $(R_5:jR_3)$, up to rotation,
is ${5 \choose j}/5$.
(There are $5 \choose j$ ways to place the $R_3$'s, but one must
divide by 5 to avoid counting rotations of the same relation.)
If $10 \le w(S) \le 12$, then write $S$ as in~(\ref{peel}).
If some $f_i$ consists of zero or one roots of unity, then the
argument of Lemma~\ref{pigeonhole} applies, and $S$ must be of the
form $(R_5:jR_3)$ with $j \le 4$,
which contradicts the last equation in the Lemma.
Otherwise the numbers of (sixth) roots of unity occurring
in $f_0,f_1,f_2,f_3,f_4$ must
be 2,2,2,2,2 or 2,2,2,2,3 or 2,2,2,3,3 or 2,2,2,2,4 in some order.
So the common value of the $f_i$ is a sum of two sixth roots of unity.
By rotating by a sixth root of unity, we may assume this value
is 0, 1, $1+\zeta_6$, or 2.
If it is 0 or 1, then the arguments in the proof of
Lemma~\ref{pigeonhole} apply.
Next assume it is $1+\zeta_6$.
The only way two sixth roots of unity can sum to $1+\zeta_6$
is if they are 1 and $\zeta_6$ in some order.
The only ways three sixth roots of unity can sum to $1+\zeta_6$
is if they are $1,1,\zeta_6^2$ or $\zeta_6,\zeta_6,\zeta_6^{-1}$.
So if the numbers of roots of unity occurring in $f_0,f_1,f_2,f_3,f_4$
are 2,2,2,2,2 or 2,2,2,2,3, then $S$ will contain $R_5$ or its rotation
by $\zeta_6$, and the same will be true for 2,2,2,3,3 unless the
two $f_i$ with three terms are $1+1+\zeta_6^2$
and $\zeta_6+\zeta_6+\zeta_6^{-1}$, in which case $S$
contains $(R_5:R_3)$.
It is impossible to write $1+\zeta_6$ as a sum of sixth roots
of unity without using 1 or $\zeta_6$, so if the numbers
are 2,2,2,2,4, then again $S$ contains $R_5$ or its rotation by $\zeta_6$.
Thus we get no new relations where the common value of
the $f_i$ is $1+\zeta_6$.
Lastly, assume this common value is $2$.
Any representation of $2$ as a sum of four or fewer sixth roots
of unity contains $1$, unless
it is $\zeta_6+\zeta_6+\zeta_6^{-1}+\zeta_6^{-1}$, so $S$ will
contain $R_5$ except possibly in the case where $f_0,f_1,f_2,f_3,f_4$
are 2,2,2,2,4 in some order, and the 4 is as above.
But in this final remaining case, $S$ contains $(R_5:R_3)$.
Thus there are no minimal relations $S$ with $p_s=5$ and $10 \le w(S) \le 12$.
\medskip
\noindent{\em Case 3: } $p_s=7$
Since $w(S) \le 12 < 2 \cdot 7$, we can apply Lemma~\ref{pigeonhole}.
Now the sum of $w(T_i)-2$ is required to be $w(S)-7$ which
is at most 5, so the $T$'s that may be used
are $R_3$, $R_5$, $(R_5:R_3)$, and the two of type $(R_5:2R_3)$,
for which weight minus 2 equals 1, 3, 4, and 5, respectively.
So the problem is reduced to listing the partitions of $w(S)-7$
into parts of size 1, 3, 4, and 5.
If all parts used are 1, then we get $(R_7:jR_3)$ with $j=w(S)-7$,
and there are ${7 \choose j}/7$ distinct relations in this class.
Otherwise exactly one part of size 3, 4, or 5 is used, and the
possibilities are as follows.
If a part of size 3 is used, we get $(R_7:R_5)$, $(R_7:R_5,R_3)$,
or $(R_7:R_5,2R_3)$, of weights 10, 11, 12 respectively.
By rotation, the $R_5$ may be assumed to share the 1 in the $R_7$,
and then there are $6 \choose i$ ways to place the $R_3$'s where $i$
is the number of $R_3$'s.
If a part of size 4 is used, we get $(R_7:(R_5:R_3))$ of weight 11
or $(R_7:(R_5:R_3),R_3)$ of weight 12.
By rotation, the $(R_5:R_3)$ may be assumed to share the 1 in the $R_7$,
but any of the six roots of unity in the $(R_5:R_3)$ may be rotated to be 1.
The $R_3$ can then overlap any of the other $6$ seventh roots of unity.
Finally, if a part of size 5 is used, we get $(R_7:(R_5:2R_3))$.
There are two different relations of type $(R_5:2R_3)$ that may be used,
and each has seven roots of unity which may be rotated to be the~1 shared
by the $R_7$, so there are 14 of these all together.
\medskip
\noindent{\em Case 4: } $p_s=11$
Applying Lemma~\ref{pigeonhole} shows that the only possibilities
are $R_{11}$ of weight 11, and $(R_{11}:R_3)$ of weight 12.
\end{pf*}
Now a general relation of weight 12 is a sum of the minimal ones of
weight up to 12, and we can classify them according to the weights
of the minimal relations, which form a partition of 12 with no parts
of size 1 or 4.
We will use the notation $(R_5:2R_3)+2R_3$, for example, to denote
a sum of three minimal relations of type $(R_5:2R_3)$, $R_3$, and $R_3$.
Table~\ref{partitions} lists the possibilities.
The parts may be rotated independently, so any category
involving more than one minimal relation contains
infinitely many relations, even up to rotation (of the entire relation).
Also, the categories are not mutually exclusive, because of the
non-uniqueness of the decomposition into minimal relations.
\begin{table}
\centerline{
\begin{tabular}{c|c}
Partition & Relation type \\ \hline \hline
12 & $(R_7:5R_3)$ \\ \cline{2-2}
& $(R_7:R_5,2R_3)$ \\ \cline{2-2}
& $(R_7:(R_5:R_3),R_3)$ \\ \cline{2-2}
& $(R_7:(R_5:2R_3))$ \\ \cline{2-2}
& $(R_{11}:R_3)$ \\ \hline
10,2 & $(R_7:3R_3)+R_2$ \\ \cline{2-2}
& $(R_7:R_5)+R_2$ \\ \hline
9,3 & $(R_5:4R_3)+R_3$ \\ \cline{2-2}
& $(R_7:2R_3)+R_3$ \\ \hline
8,2,2 & $(R_5:3R_3)+2R_2$ \\ \cline{2-2}
& $(R_7:R_3)+2R_2$ \\ \hline
\end{tabular}
\hfil
\begin{tabular}{c|c}
Partition & Relation type \\ \hline \hline
7,5 & $(R_5:2R_3)+R_5$ \\ \cline{2-2}
& $R_7+R_5$ \\ \hline
7,3,2 & $(R_5:2R_3)+R_3+R_2$ \\ \cline{2-2}
& $R_7+R_3+R_2$ \\ \hline
6,6 & $2(R_5:R_3)$ \\ \hline
6,3,3 & $(R_5:R_3)+2R_3$ \\ \hline
6,2,2,2 & $(R_5:R_3)+3R_2$ \\ \hline
5,5,2 & $2R_5+R_2$ \\ \hline
5,3,2,2 & $R_5+R_3+2R_2$ \\ \hline
3,3,3,3 & $4R_3$ \\ \hline
3,3,2,2,2 & $2R_3+3R_2$ \\ \hline
2,2,2,2,2,2 & $6R_2$ \\ \hline
\end{tabular}
}
\vskip12pt
\caption{The types of relations of weight 12.}
\label{partitions}
\end{table}
\section{Solutions to the trigonometric equation}
\label{solutions}
Here we use the classification of the previous section to give a complete
listing of the solutions to the trigonometric equation~(\ref{product2}).
There are some obvious solutions to~(\ref{product2}), namely those in
which $U,V,W$ are arbitary positive rational numbers with sum $1/2$,
and $X,Y,Z$ are a permutation of $U,V,W$.
We will call these the trivial solutions, even though the three-diagonal
intersections they give rise to can look surprising.
See Figure~\ref{16gon} for an example on the 16-gon.
\begin{figure}
\centerline{\psfig{file=fig3.ps,width=3in}}
\caption{A surprising trivial solution for the 16-gon. The intersection
point does not lie on any of the 16 lines of symmetry of the
16-gon.\null\hfill\break\null}
\label{16gon}
\end{figure}
The twelve roots of unity occurring in~(\ref{thetwelve}) are not arbitrary;
therefore we must go through Table~\ref{partitions} to see which
relations are of the correct form, i.e., expressible as a sum of six roots
of unity and their inverses, where the product of the six is -1.
First let us prove a few lemmas that will greatly reduce the number of cases.
\begin{lemma}
Let $S$ be a relation of weight $k \le 12$.
Suppose $S$ is stable under complex conjugation (i.e., under
$\zeta \mapsto \zeta^{-1}$).
Then $S$ has a complex conjugation-stable decomposition
into minimal relations; i.e., each minimal relation occurring is itself
stable under complex conjugation, or can be paired with another minimal
relation which is its complex conjugate.
\end{lemma}
\begin{pf}
We will use induction on $k$.
If $S$ is minimal, there is nothing to prove.
Otherwise let $T$ be a (minimal) subrelation of $S$ of minimal weight,
so $T$ is of weight at most 6.
The complex conjugate $\overline{T}$ of $T$ is another minimal relation
in $S$.
If they do not intersect, then we take the decomposition of $S$ into $T$,
$\overline{T}$, and a decomposition of $S \setminus (T \cup \overline{T})$
given by the inductive hypothesis.
If they do overlap and the weight of $T$ is at most 5, then $T=R_p$ for
some prime $p$, and the fact that $T$ intersects $\overline{T}$ implies
that $T=\overline{T}$, and we get the result by applying the inductive
hypothesis to $S \setminus T$.
The only remaining case is where $S$ is of type $2(R_5:R_3)$.
If the two $(R_5:R_3)$'s are not conjugate to each other, then for each
there is a root of unity $\zeta$ such that $\zeta$ and $\zeta^{-1}$ occur
in that (rotation of) $(R_5:R_3)$.
The quotient $\zeta^2$ is then a $30$-th root of unity, so $\zeta$
itself is a $60$-th root of unity.
Thus each $(R_5:R_3)$ is a rotation of the ``standard'' $(R_5:R_3)$
as in~(\ref{r5r3}) by a $60$-th root of unity, and we let Mathematica
check the $60^2$ possibilities.
\end{pf}
We do not know if the preceding lemma holds for relations of weight
greater than 12.
\begin{lemma}
Let $S$ be a minimal relation of type $(R_p:T_1,\ldots,T_j)$, $p \ge 5$,
where the $T_i$ involve roots of unity of order prime to $p$, and $j<p$.
If $S$ is stable under complex conjugation, then the particular rotation
of $R_p$ from which the $T_i$ were ``subtracted'' is also stable
(and hence so is the collection of the relations subtracted).
\end{lemma}
\begin{pf}
Let $\ell$ be the product of the orders of the roots of unity in
all the $T_i$.
The elements of $S$ in the original $R_p$ can be characterized
as those terms of $S$ that are unique in their coset of $\mu_\ell$
(the $\ell$-th roots of unity), and this condition is stable under
complex conjugation, so the set of terms of the $R_p$ that were not
subtracted is stable.
Since $j<p$, we can pick one such term $\zeta$.
Then the quotient $\zeta/\zeta^{-1}$ is a $p$-th root of unity,
so $\zeta$ is a $2p$-th root of unity, and hence the $R_p$
containing it is stable.
\end{pf}
\begin{cor}
A relation of type $(R_7:(R_5:R_3),R_3)$ cannot be stable under
complex conjugation.
\end{cor}
Even with these restrictions, a very large number of cases remain,
so we perform the calculation using Mathematica.
Each entry of Table~\ref{partitions} represents a finite number of
linearly parameterized (in the exponents) families of relations of weight 12.
For each parameterized family, we check to see what additional
constraints must be put on the parameters for the relation to be
of the form of~(\ref{thetwelve}).
Next, for each parameterized family of solutions to~(\ref{thetwelve}),
we calculate the corresponding $U,V,W,X,Y,Z$ and throw away solutions
in which some of these are nonpositive.
Finally, we sort $U,V,W$ and $X,Y,Z$ and interchange the two triples
if $U>X$, in order to count the solutions only up to symmetry.
The results of this computation are recorded in the following theorem.
\begin{theorem}
\label{thesolutions}
The positive rational solutions to~(\ref{product2}), up to symmetry,
can be classified as follows:
\begin{enumerate}
\item The trivial solutions,
which arise from relations of type $6R_2$.
\item Four one-parameter families of solutions,
listed in Table~\ref{families}.
The first arises from relations of type $4R_3$,
and the other three arise from relations of type $2R_3+3R_2$.
\item Sixty-five ``sporadic'' solutions, listed in
Table~\ref{sporadics}, which arise from the other types
of weight 12 relations listed in Table~\ref{partitions}.
\end{enumerate}
The only duplications in this list are that the second family of
Table~\ref{families} gives a trivial solution for $t=1/12$,
the first and fourth families of Table~\ref{families} give the same
solution when $t=1/18$ in both, and the second and fourth families
of Table~\ref{families} give the same solution when $t=1/24$ in both.
\end{theorem}
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c||c|c|c||c|}
$U$ & $V$ & $W$ & $X$ & $Y$ & $Z$ & Range \\ \hline \hline
$1/6$ & $t$ & $1/3-2t$ & $1/3+t$ & $t$ & $1/6-t$ & $0<t<1/6$ \\ \hline
$1/6$ & $1/2-3t$ & $t$ & $1/6-t$ & $2t$ & $1/6+t$ & $0<t<1/6$ \\ \hline
$1/6$ & $1/6-2t$ & $2t$ & $1/6-2t$ & $t$ & $1/2+t$ & $0<t<1/12$ \\ \hline
$1/3-4t$ & $t$ & $1/3+t$ & $1/6-2t$ & $3t$ & $1/6+t$ & $0<t<1/12$ \\ \hline
\end{tabular}
\end{center}
\caption{The nontrivial infinite families of solutions to~(\ref{product2}).}
\label{families}
\end{table}
\begin{table}
{\tiny
\begin{center}
\begin{tabular}{c||c|c|c||c|c|c||c}
Denominator & $U$ & $V$ & $W$ & $X$ & $Y$ & $Z$
& Relation type \\ \hline \hline
30 & $1/10$ & $2/15$ & $3/10$ & $2/15$ & $1/6$ & $1/6$
& $2(R_5:R_3)$ \\ \cline{2-7}
& $1/15$ & $1/15$ & $7/15$ & $1/15$ & $1/10$ & $7/30$ & \\ \cline{2-7}
& $1/30$ & $7/30$ & $4/15$ & $1/15$ & $1/10$ & $3/10$ & \\ \cline{2-7}
& $1/30$ & $1/10$ & $7/15$ & $1/15$ & $1/15$ & $4/15$ & \\ \cline{2-7}
& $1/30$ & $1/15$ & $19/30$ & $1/15$ & $1/10$ & $1/10$ & \\ \cline{2-8}
& $1/15$ & $1/6$ & $4/15$ & $1/10$ & $1/10$ & $3/10$
& $(R_5:R_3)+2R_3$ \\ \cline{2-7}
& $1/15$ & $2/15$ & $11/30$ & $1/10$ & $1/6$ & $1/6$ & \\ \cline{2-7}
& $1/30$ & $1/6$ & $13/30$ & $1/10$ & $2/15$ & $2/15$ & \\ \cline{2-7}
& $1/30$ & $1/30$ & $7/10$ & $1/30$ & $1/15$ & $2/15$ & \\ \cline{2-8}
& $1/30$ & $7/30$ & $3/10$ & $1/15$ & $2/15$ & $7/30$
& $R_5+R_3+2R_2$ \\ \cline{2-7}
& $1/30$ & $1/6$ & $11/30$ & $1/15$ & $1/10$ & $4/15$ & \\ \cline{2-7}
& $1/30$ & $1/10$ & $13/30$ & $1/30$ & $2/15$ & $4/15$ & \\ \cline{2-7}
& $1/30$ & $1/15$ & $8/15$ & $1/30$ & $1/10$ & $7/30$ & \\ \hline
42 & $1/14$ & $5/42$ & $5/14$ & $2/21$ & $5/42$ & $5/21$
& $(R_7:5R_3)$ \\ \cline{2-7}
& $1/21$ & $4/21$ & $13/42$ & $1/14$ & $1/6$ & $3/14$ & \\ \cline{2-7}
& $1/42$ & $3/14$ & $5/14$ & $1/21$ & $1/6$ & $4/21$ & \\ \cline{2-7}
& $1/42$ & $1/6$ & $19/42$ & $1/14$ & $2/21$ & $4/21$ & \\ \cline{2-7}
& $1/42$ & $1/6$ & $13/42$ & $1/21$ & $1/14$ & $8/21$ & \\ \cline{2-7}
& $1/42$ & $1/21$ & $13/21$ & $1/42$ & $1/14$ & $3/14$ & \\ \hline
60 & $1/20$ & $1/12$ & $29/60$ & $1/15$ & $1/10$ & $13/60$
& $2(R_5:R_3)$ \\ \cline{2-7}
& $1/20$ & $1/12$ & $9/20$ & $1/15$ & $1/12$ & $4/15$ & \\ \cline{2-7}
& $1/20$ & $1/12$ & $5/12$ & $1/20$ & $1/10$ & $3/10$ & \\ \cline{2-7}
& $1/60$ & $4/15$ & $3/10$ & $1/20$ & $1/12$ & $17/60$ & \\ \cline{2-7}
& $1/60$ & $13/60$ & $9/20$ & $1/12$ & $1/10$ & $2/15$ & \\ \cline{2-7}
& $1/60$ & $13/60$ & $5/12$ & $1/20$ & $2/15$ & $1/6$ & \\ \cline{2-8}
& $1/12$ & $1/6$ & $17/60$ & $2/15$ & $3/20$ & $11/60$
& $(R_5:3R_3)+2R_2$ \\ \cline{2-7}
& $1/12$ & $2/15$ & $19/60$ & $1/10$ & $3/20$ & $13/60$ & \\ \cline{2-7}
& $1/15$ & $11/60$ & $13/60$ & $1/12$ & $1/10$ & $7/20$ & \\ \cline{2-7}
& $1/20$ & $11/60$ & $3/10$ & $1/12$ & $7/60$ & $4/15$ & \\ \cline{2-7}
& $1/20$ & $1/10$ & $23/60$ & $1/15$ & $1/12$ & $19/60$ & \\ \cline{2-7}
& $1/30$ & $7/60$ & $19/60$ & $1/20$ & $1/15$ & $5/12$ & \\ \cline{2-7}
& $1/30$ & $1/12$ & $7/12$ & $1/15$ & $1/10$ & $2/15$ & \\ \cline{2-7}
& $1/30$ & $1/20$ & $11/20$ & $1/30$ & $1/15$ & $4/15$ & \\ \cline{2-7}
& $1/60$ & $3/10$ & $7/20$ & $1/12$ & $7/60$ & $2/15$ & \\ \cline{2-7}
& $1/60$ & $4/15$ & $23/60$ & $1/12$ & $1/10$ & $3/20$ & \\ \cline{2-7}
& $1/60$ & $7/30$ & $5/12$ & $1/15$ & $7/60$ & $3/20$ & \\ \cline{2-7}
& $1/60$ & $13/60$ & $11/30$ & $1/20$ & $1/12$ & $4/15$ & \\ \cline{2-7}
& $1/60$ & $1/6$ & $31/60$ & $1/15$ & $1/10$ & $2/15$ & \\ \cline{2-7}
& $1/60$ & $1/6$ & $5/12$ & $1/20$ & $1/15$ & $17/60$ & \\ \cline{2-7}
& $1/60$ & $2/15$ & $9/20$ & $1/30$ & $1/12$ & $17/60$ & \\ \cline{2-7}
& $1/60$ & $1/10$ & $31/60$ & $1/30$ & $1/15$ & $4/15$ & \\ \hline
84 & $1/12$ & $3/14$ & $19/84$ & $11/84$ & $13/84$ & $4/21$
& $(R_7:R_3)+2R_2$ \\ \cline{2-7}
& $1/14$ & $11/84$ & $23/84$ & $1/12$ & $2/21$ & $29/84$ & \\ \cline{2-7}
& $1/21$ & $13/84$ & $23/84$ & $1/14$ & $1/12$ & $31/84$ & \\ \cline{2-7}
& $1/42$ & $1/12$ & $7/12$ & $1/21$ & $1/14$ & $4/21$ & \\ \cline{2-7}
& $1/84$ & $25/84$ & $5/14$ & $5/84$ & $1/12$ & $4/21$ & \\ \cline{2-7}
& $1/84$ & $5/21$ & $5/12$ & $5/84$ & $1/14$ & $17/84$ & \\ \cline{2-7}
& $1/84$ & $3/14$ & $37/84$ & $1/21$ & $1/12$ & $17/84$ & \\ \cline{2-7}
& $1/84$ & $1/6$ & $43/84$ & $1/21$ & $1/14$ & $4/21$ & \\ \hline
90 & $1/18$ & $13/90$ & $7/18$ & $11/90$ & $2/15$ & $7/45$
& $(R_5:R_3)+2R_3$ \\ \cline{2-7}
& $1/45$ & $19/90$ & $16/45$ & $1/18$ & $1/10$ & $23/90$ & \\ \cline{2-7}
& $1/90$ & $23/90$ & $31/90$ & $2/45$ & $1/15$ & $5/18$ & \\ \cline{2-7}
& $1/90$ & $17/90$ & $47/90$ & $1/18$ & $4/45$ & $2/15$ & \\ \hline
120 & $13/120$ & $3/20$ & $31/120$ & $2/15$ & $19/120$ & $23/120$ &
$(R_5:R_3)+3R_2$ \\ \cline{2-7}
& $1/12$ & $19/120$ & $29/120$ & $1/10$ & $13/120$ & $37/120$ & \\ \cline{2-7}
& $1/20$ & $23/120$ & $29/120$ & $1/15$ & $13/120$ & $41/120$ & \\ \cline{2-7}
& $1/60$ & $13/120$ & $73/120$ & $1/20$ & $1/12$ & $2/15$ & \\ \cline{2-7}
& $1/120$ & $7/20$ & $43/120$ & $7/120$ & $11/120$ & $2/15$ & \\ \cline{2-7}
& $1/120$ & $3/10$ & $49/120$ & $7/120$ & $1/12$ & $17/120$ & \\ \cline{2-7}
& $1/120$ & $4/15$ & $53/120$ & $1/20$ & $11/120$ & $17/120$ & \\ \cline{2-7}
& $1/120$ & $13/60$ & $61/120$ & $1/20$ & $1/12$ & $2/15$ & \\ \hline
210 & $1/15$ & $41/210$ & $8/35$ & $1/14$ & $31/210$ & $61/210$
& $(R_7:(R_5:2R_3))$ \\ \cline{2-7}
& $13/210$ & $1/10$ & $83/210$ & $1/14$ & $4/35$ & $9/35$ & \\ \cline{2-7}
& $1/35$ & $2/15$ & $97/210$ & $1/14$ & $17/210$ & $47/210$ & \\ \cline{2-7}
& $1/210$ & $3/14$ & $121/210$ & $11/210$ & $1/15$ & $3/35$ & \\ \hline
\end{tabular}
\end{center}
}
\caption{The 65 sporadic solutions to~(\ref{product2}).}
\label{sporadics}
\end{table}
Some explanation of the tables is in order.
The last column of Table~\ref{families} gives the allowable range
for the rational parameter $t$.
The entries of Table~\ref{sporadics} are sorted according to the least
common denominator of $U,V,W,X,Y,Z$, which is also the least $n$
for which diagonals of a regular $n$-gon can create arcs of the
corresponding lengths.
The relation type from which each solution derives is also given.
The reason 11 does not appear in the least common denominator for
any sporadic solution is that the relation $(R_{11}:R_3)$
cannot be put in the form of~(\ref{thetwelve}) with the $\alpha_j$
summing to 1, and hence leads to no solutions of~(\ref{product2}).
(Several other types of relations also give rise to no solutions.)
Tables~\ref{families} and~\ref{sporadics} are the same as Bol's tables
at the bottom of page~40 and on page~41 of~\cite{bol},
in a slightly different format.
\medskip
The arcs cut by diagonals of a regular $n$-gon have lengths which are
multiples of $2 \pi/n$, so $U$, $V$, $W$, $X$, $Y$ and $Z$ corresponding
to any configuration of three diagonals meeting must be multiples of $1/n$.
With this additional restriction, trivial solutions to~(\ref{product2})
occur only when $n$ is even (and at least 6).
Solutions within the infinite families of Table~\ref{families} occur
when $n$ is a multiple of 6 (and at least 12), and there $t$ must be
a multiple of $1/n$.
Sporadic solutions with least common denominator $d$ occur if and only
if $n$ is a multiple of $d$.
\section{Intersections of more than three diagonals}
\label{more}
Now that we know the configurations of three diagonals meeting,
we can check how they overlap to produce configurations of more
than three diagonals meeting.
We will disregard configurations in which the intersection point
is the center of the $n$-gon, since these are easily described:
there are exactly $n/2$ diagonals (diameters) through the center
when $n$ is even, and none otherwise.
When $k$ diagonals meet, they form $2k$ arcs, whose lengths we will
measure as a fraction of the whole circumference (so they will be
multiples of $1/n$) and list in counterclockwise order.
(Warning: this is different from the order used in Tables \ref{families}
and~\ref{sporadics}.)
The least common denominator of the numbers in this list will be
called the denominator of the configuration.
It is the least $n$ for which the configuration can be realized as
diagonals of a regular $n$-gon.
\begin{lemma}
\label{denominator}
If a configuration of $k \ge 2$ diagonals meeting at an interior point
other than the center has denominator dividing $d$, then any configuration
of diagonals meeting at that point has denominator dividing $\operatorname{LCM}(2d,3)$.
\end{lemma}
\begin{pf}
We may assume $k=2$.
Any other configuration of diagonals through the intersection point is
contained in the union of configurations obtained by adding one
diagonal to the original two, so we may assume the final configuration
consists of three diagonals, two of which were the original two.
Now we need only go through our list of three-diagonal intersections.
It can be checked (using Mathematica) that removing any diagonal from
a sporadic configuration of three intersecting diagonals yields a
configuration whose denominator is the same or half as much, except
that it is possible that removing a diagonal from a three-diagonal
configuration of denominator 210 or 60 yields one of denominator 70 or 20,
respectively, which proves the desired result for these cases.
The additive group generated by $1/6$ and the normalized arc lengths
of a configuration obtained by removing a diagonal from a configuration
corresponding to one of the families of Table~\ref{families}
contains $2t$ where $t$ is the parameter, (as can be verified using
Mathematica again), which means that adding that third diagonal
can at most double the denominator (and throw in a factor of 3,
if it isn't already there).
Similarly, it is easily checked (even by hand), that the subgroup
generated by the normalized arc lengths of a configuration obtained
by removing one of the three diagonals of a configuration corresponding
to a trivial solution to~(\ref{product2}) but with intersection point
not the center, contains twice the arc lengths of the original configuration.
\end{pf}
\begin{cor}
\label{partsporadic}
If a configuration of three or more diagonals meeting includes three
forming a sporadic configuration, then its
denominator is 30, 42, 60, 84, 90, 120, 168, 180, 210, 240, or 420.
\end{cor}
\begin{pf}
Combine the lemma with the list of denominators of sporadic configurations
listed in Table~\ref{sporadics}.
\end{pf}
For $k \ge 4$, a list of $2k$ positive rational numbers summing to~1
arises this way if and only if the lists of length $2k-2$ which would
arise by removing the first or second diagonal actually
correspond to $k-1$ intersecting diagonals.
Suppose $k=4$.
If we specify the sporadic configuration or parameterized family
of configurations that arise when we remove the first or second diagonal,
we get a set of linear conditions on the eight arc lengths.
Corollary~\ref{partsporadic} tells us that we get a configuration with
denominator among 30, 42, 60, 84, 90, 120, 168, 180, 210, 240, and 420,
if one of these two is sporadic.
Using Mathematica to perform this computation for the rest of
possibilities in Theorem~\ref{thesolutions} shows that the other
four-diagonal configurations, up to rotation and reflection,
fall into 12 one-parameter families, which are listed in
Table~\ref{theeights} by the eight normalized arc lengths
and the range for the parameter $t$, with a finite number of exceptions
of denominators among 12, 18, 24, 30, 36, 42, 48, 60, 84, and 120.
\begin{table}
\centerline{\small
\begin{tabular}{|c|c|c|c|c|c|c|c||c|}
& & & & & & & & Range \\ \hline \hline
$t$ & $t$ & $t$ & $1/6-2t$
& $1/6$ & $1/3+t$ & $1/6$ & $1/6-2t$ & $0<t<1/12$ \\ \hline
$t$ & $1/6-t$ & $1/6-t$ & $1/6-t$
& $t$ & $1/6$ & $1/6+t$ & $1/6$ & $0<t<1/6$ \\ \hline
$1/6-4t$ & $2t$ & $t$ & $3t$
& $1/6-4t$ & $1/6$ & $1/6+t$ & $1/3+t$ & $0<t<1/24$ \\ \hline
$2t$ & $1/2-t$ & $2t$ & $1/6-2t$
& $t$ & $1/6-t$ & $t$ & $1/6-2t$ & $0<t<1/12$ \\ \hline
$1/3-4t$ & $1/6+t$ & $1/2-3t$ & $-1/6+4t$
& $1/6-2t$ & $t$ & $1/6-t$ & $-1/6+4t$ & $1/24<t<1/12$ \\ \hline
$2t$ & $t$ & $3t$ & $1/6-2t$
& $1/6$ & $1/6-t$ & $1/3-t$ & $1/6-2t$ & $0<t<1/12$ \\ \hline
$t$ & $t$ & $2t$ & $1/3-t$
& $1/6$ & $1/6-t$ & $1/6-t$ & $1/6-t$ & $0<t<1/6$ \\ \hline
$1/3-4t$ & $1/6$ & $t$ & $t$
& $1/6-2t$ & $1/3-2t$ & $3t$ & $3t$ & $0<t<1/12$ \\ \hline
$2t$ & $1/3-2t$ & $1/6-t$ & $1/6-t$
& $1/6$ & $1/6$ & $t$ & $t$ & $0<t<1/6$ \\ \hline
$1/3-4t$ & $2t$ & $t$ & $t$
& $1/6-2t$ & $1/6$ & $1/6+t$ & $1/6+t$ & $0<t<1/12$ \\ \hline
$1/3-4t$ & $2t$ & $1/6-t$ & $t$
& $1/6-2t$ & $2t$ & $1/3-t$ & $3t$ & $0<t<1/12$ \\ \hline
$2t$ & $1/6-t$ & $t$ & $1/6-t$
& $t$ & $1/6-t$ & $2t$ & $1/2-3t$ & $0<t<1/6$ \\ \hline
\end{tabular}
}
\vskip10pt
\caption{The one-parameter families of four-diagonal configurations.}
\label{theeights}
\end{table}
We will use a similar argument when $k=5$.
Any five-diagonal configuration containing a sporadic three-diagonal
configuration will again have denominator
among 30, 42, 60, 84, 90, 120, 168, 180, 210, 240, and 420.
Any other five-diagonal configuration containing one of the exceptional
four-diagonal configurations will have denominator
among 12, 18, 24, 30, 36, 42, 48, 60, 72, 84, 96, 120, 168, and 240,
by Lemma~\ref{denominator}.
Finally, another Mathematica computation shows that the one-parameter
families of four-diagonal configurations overlap to produce the
one-parameter families listed (up to rotation and reflection)
in Table~\ref{thetens}, and a finite number of exceptions of
denominators among 18, 24, and 30.
\begin{table}
\centerline{\small
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c||c|}
& & & & & & & & & & Range \\ \hline \hline
$t$ & $2t$ & $1/6-2t$ & $1/6$ & $1/6-t$
& $1/6-t$ & $1/6$ & $1/6-2t$ & $2t$ & $t$ & $0<t<1/12$ \\ \hline
$t$ & $2t$ & $1/6-4t$ & $1/6$ & $1/6+t$
& $1/6+t$ & $1/6$ & $1/6-4t$ & $2t$ & $t$ & $0<t<1/24$ \\ \hline
$t$ & $1/6-2t$ & $-1/6+4t$ & $1/3-4t$ & $1/6+t$ & $1/6+t$
& $1/3-4t$ & $-1/6+4t$ & $1/6-2t$ & $t$ & $1/24<t<1/12$ \\ \hline
$t$ & $1/6-2t$ & $2t$ & $1/3-4t$ & $3t$ & $3t$
& $1/3-4t$ & $2t$ & $1/6-2t$ & $t$ & $0<t<1/12$ \\ \hline
\end{tabular}
}
\vskip10pt
\caption{The one-parameter families of five-diagonal configurations.}
\label{thetens}
\end{table}
For $k=6$, any six-diagonal configuration containing a sporadic
three-diagonal configuration will again have denominator
among 30, 42, 60, 84, 90, 120, 168, 180, 210, 240, and 420.
Any six-diagonal configuration containing one of the exceptional
four-diagonal configurations will have denominator
among 12, 18, 24, 30, 36, 42, 48, 60, 72, 84, 96, 120, 168, and 240.
Any six-diagonal configuration containing one of the
exceptional five-diagonal configurations will have denominator
among 18, 24, 30, 36, 48, and 60.
Another Mathematica computation shows that the one-parameter families
of five-diagonal configurations cannot combine to give a
six-diagonal configuration.
Finally for $k \ge 7$, any $k$-diagonal configuration must contain
an exceptional configuration of 3, 4, or 5 diagonals, and hence by
Lemma~\ref{denominator} has denominator
among 12, 18, 24, 30, 36, 42, 48, 60, 72, 84, 90, 96,
120, 168, 180, 210, 240, and 420.
We summarize the results of this section in the following.
\begin{prop}
\label{configurations}
The configurations of $k \ge 4$ diagonals meeting at a point
not the center, up to rotation and reflection, fall into the
one-parameter families listed in Tables \ref{theeights} and~\ref{thetens},
with finitely many exceptions (for fixed $k$) of denominators
among 12, 18, 24, 30, 36, 42, 48, 60, 72, 84, 90, 96, 120, 168,
180, 210, 240, and 420.
\end{prop}
In fact, many of the numbers listed in the proposition do not
actually occur as denominators of exceptional configurations.
For example, it will turn out that the only denominator greater
than 120 that occurs is 210.
\section{The formula for intersection points}
\label{points}
Let $a_k(n)$ denote the number of points inside the regular $n$-gon
other than the center where exactly $k$ lines meet.
Let $b_k(n)$ denote the number of $k$-tuples of diagonals which meet
at a point inside the $n$-gon other than the center.
Each interior point at which exactly $m$ diagonals meet gives rise
to $m \choose k$ such $k$-tuples, so we have the relationship
\begin{equation}
\label{relationship}
b_k(n) = \sum_{m \ge k} {m \choose k} a_m(n)
\end{equation}
Since every four distinct vertices of the $n$-gon determine one pair
of diagonals which intersect inside, the number of such pairs is
exactly $n \choose 4$, but if $n$ is even, then ${n/2} \choose 2$
of these are pairs which meet at the center, so
\begin{equation}
\label{b2}
b_2(n)={n \choose 4} - {{n/2} \choose 2} \delta_2(n).
\end{equation}
(Recall that $\delta_m(n)$ is defined to be 1 if $n$ is a multiple of $m$,
and 0 otherwise.)
We will use the results of the previous two sections to deduce the
form of $b_k(n)$ and then the form of $a_k(n)$.
To avoid having to repeat the following, let us make a definition.
\begin{defn}
A function on integers $n \ge 3$ will be called {\em tame}
if it is a linear combination (with rational coefficients)
of the functions $n^3$, $n^2$, $n$, $1$, $n^2 \delta_2(n)$,
$n \delta_2(n)$, $\delta_2(n)$, $\delta_4(n)$, $n \delta_6(n)$,
$\delta_6(n)$, $\delta_{12}(n)$, $\delta_{18}(n)$, $\delta_{24}(n)$,
$\delta_{24}(n-6)$, $\delta_{30}(n)$, $\delta_{36}(n)$, $\delta_{42}(n)$,
$\delta_{48}(n)$, $\delta_{60}(n)$, $\delta_{72}(n)$, $\delta_{84}(n)$,
$\delta_{90}(n)$, $\delta_{96}(n)$, $\delta_{120}(n)$, $\delta_{168}(n)$,
$\delta_{180}(n)$, $\delta_{210}(n)$, and $\delta_{420}(n)$.
\end{defn}
\begin{prop}
\label{form}
For each $k \ge 2$, the function $b_k(n)/n$ on integers $n \ge 3$ is tame.
\end{prop}
\begin{pf}
The case $k=2$ is handled by~(\ref{b2}), so assume $k \ge 3$.
Each list of $2k$ normalized arc lengths as in Section~\ref{more}
corresponding to a configuration of $k$ diagonals meeting at a point
other than the center, considered up to rotation (but not reflection),
contributes $n$ to $b_k(n)$.
(There are $n$ places to start measuring the arcs from, and these $n$
configurations are distinct, because the corresponding intersection points
differ by rotations of multiples of $2 \pi/n$, and by assumption they are
not at the center.)
So $b_k(n)/n$ counts such lists.
Suppose $k=3$.
When $n$ is even, the family of trivial solutions to the trigonometric
equation~(\ref{product2}) has $U=a/n$, $V=b/n$, $W=c/n$,
where $a$, $b$, and $c$ are positive integers with sum $n/2$,
and $X$, $Y$, and $Z$ are some permutation of $U$, $V$, $W$.
Each permutation gives rise to a two-parameter family of six-long lists
of arc lengths, and the number of lists within each family is the number
of partitions of $n/2$ into three positive parts, which is a quadratic
polynomial in $n$.
Similarly each family of solutions in Table~\ref{families} gives rise
to a number of one-parameter families of lists, when $n$ is a multiple of 6,
each containing $\lceil n/6 \rceil - 1$ or $\lceil n/12 \rceil - 1$ lists.
These functions of $n$ (extended to be 0 when 6 does not divide $n$)
are expressible as a linear combination of $n \delta_6(n)$, $\delta_6(n)$,
and $\delta_{12}(n)$.
Finally the sporadic solutions to~\ref{product2} give rise to a
finite number of lists, having denominators among 30, 42, 60, 84, 90, 120,
and 210, so their contribution to $b_3(n)/n$ is a linear combination
of $\delta_{30}(n), \ldots, \delta_{210}(n)$.
But these families of lists overlap, so we must use the Principle of
Inclusion-Exclusion to count them properly.
To show that the result is a tame function, it suffices to show that
the number of lists in any intersection of these families is a tame function.
When two of the trivial families overlap but do not coincide, they
overlap where two of the $a$, $b$, and $c$ above are equal, and the
corresponding lists lie in one of the one-parameter families
$(t,t,t,t,1/2-2t,1/2-2t)$ or $(t,t,t,1/2-2t,t,1/2-2t)$ (with $0<t<1/4$),
each of which contain $\lceil n/4 \rceil -1$ lists (for $n$ even).
This function of $n$ is a combination of $n \delta_2(n)$, $\delta_2(n)$,
and $\delta_4(n)$, hence it is tame.
Any other intersection of the infinite families must contain the
intersection of two one-parameter families which are among the two
above or arise from Table~\ref{families}, and a Mathematica
computation shows that such an intersection consists of at most a
single list of denominator among 6, 12, 18, 24, and 30.
And, of course, any intersection involving a single sporadic list,
can contain at most that sporadic list.
Thus the number of lists within any intersection is a tame function
of $n$.
Finally we must delete the lists which correspond to configurations
of diagonals meeting at the center.
These are the lists within the trivial two-parameter family
$(t,u,1/2-t-u,t,u,1/2-t-u)$, so their number is also a tame function
of $n$, by the Principle of Inclusion-Exclusion again.
Thus $b_3(n)/n$ is tame.
Next suppose $k=4$.
The number of lists within each family listed in Table~\ref{theeights},
or the reflection of such a family, is (when $n$ is divisible by 6)
the number of multiples of $1/n$ strictly between $\alpha$ and $\beta$,
where the range for the parameter $t$ is $\alpha<t<\beta$.
This number is $\lceil \beta n \rceil - 1 - \lfloor \alpha n \rfloor$.
Since the table shows that $\alpha$ and $\beta$ are always multiples
of $1/24$, this function of $n$ is expressible as a combination
of $n \delta_6(n)$ and a function on multiples of 6 depending only
on $n \bmod 24$, and the latter can be written as a combination
of $\delta_6(n)$, $\delta_{12}(n)$, $\delta_{24}(n)$,
and $\delta_{24}(n-6)$, so it is tame.
Mathematica shows that when two of these families are not the same,
they intersect in at most a single list of denominator
among 6, 12, 18, and 24.
So these and the exceptions of Proposition~\ref{configurations}
can be counted by a tame function.
Thus, again by the Principle of Inclusion-Exclusion, $b_4(n)/n$ is tame.
The proof for $k=5$ is identical to that of $k=4$, using
Table~\ref{thetens} instead of Table~\ref{theeights},
and using another Mathematica computation which shows that
the intersections of two one-parameter families of lists consist
of at most a single list of denominator 24.
The proof for $k \ge 6$ is even simpler, because then there are
only the exceptional lists.
By Proposition~\ref{configurations}, $b_k(n)/n$ is a linear
combination of $\delta_m(n)$ where $m$ ranges over the possible
denominators of exceptional lists listed in the proposition, so it is tame.
\end{pf}
\begin{lemma}
\label{determined}
A tame function is determined by its values
at $n=$ 3, 4, 5, 6, 7, 8, 9, 10, 12, 18, 24, 30, 36, 42, 48, 54, 60,
66, 72, 84, 90, 96, 120, 168, 180, 210, and 420.
\end{lemma}
\begin{pf}
By linearity, it suffices to show that if a tame function $f$
is zero at those values, then $f$ is the zero linear combination
of the functions in the definition of a tame function.
The vanishing at $n=3$, 5, 7, and 9 forces the coefficients
of $n^3$, $n^2$, $n$, and $1$ to vanish, by Lagrange interpolation.
Then comparing the values at $n=4$ and $n=10$ shows that the
coefficient of $\delta_4(n)$ is zero.
The vanishing at $n=4$, 8, and 10 forces the coefficients
of $n^2 \delta_2(n)$, $n \delta_2(n)$, and $\delta_2(n)$ to vanish.
Comparing the values at $n=6$ and $n=54$ shows that the
coefficient of $n \delta_6$ is zero.
Comparing the values at $n=6$ and $n=66$ shows that the
coefficient of $\delta_24(n-6)$ is zero.
At this point, we know that $f(n)$ is a combination of $\delta_m(n)$,
for $m=6$, 12, 18, 24, 30, 36, 42, 48, 60, 72, 84, 90, 96, 120,
168, 180, 210, and 420.
For each $m$ in turn, $f(m)=0$ now implies that the coefficient
of $\delta_m(n)$ is zero.
\end{pf}
\begin{pf*}{Proof of Theorem~\ref{countintersections}}
Computation (see the appendix) shows that the tame function $b_8(n)/n$
vanishes at all the numbers listed in Lemma~\ref{determined}.
Hence by that lemma, $b_8(n)=0$ for all $n$.
Thus by~(\ref{relationship}), $a_k(n)$ and $b_k(n)$ are identically zero
for all $k \ge 8$ as well.
By reverse induction on $k$, we can invert~(\ref{relationship})
to express $a_k(n)$ as a linear combination of $b_m(n)$ with $m \ge k$.
Hence $a_k(n)/n$ is tame as well for each $k \ge 2$.
Computation shows that the equations
\begin{eqnarray*}
a_2(n)/n & = & (n^3 - 6n^2 + 11n - 6)/24
+ (-5 n^2 + 46 n - 72)/16 \cdot \delta_2(n) \\
& & \mbox{} - 9/4 \cdot \delta_4(n)
+ (-19 n + 110)/2 \cdot \delta_6(n)
+ 54 \cdot \delta_{12}(n) + 84 \cdot \delta_{18}(n) \\
& & \mbox{} + 50 \cdot \delta_{24}(n)
- 24 \cdot \delta_{30}(n) - 100 \cdot \delta_{42}(n)
- 432 \cdot \delta_{60}(n) \\
& & \mbox{} -204 \cdot \delta_{84}(n)
- 144 \cdot \delta_{90}(n) - 204 \cdot \delta_{120}(n)
- 144 \cdot \delta_{210}(n) \\
a_3(n)/n & = & (5 n^2 - 48 n + 76)/48 \cdot \delta_2(n)
+ 3/4 \cdot \delta_4(n) + (7n - 38)/6 \cdot \delta_6(n)\\
& & \mbox{} - 8 \cdot \delta_{12}(n)
- 20 \cdot \delta_{18}(n) - 16 \cdot \delta_{24}(n)
- 19 \cdot \delta_{30}(n) + 8 \cdot \delta_{42}(n) \\
& & \mbox{} + 68 \cdot \delta_{60}(n)
+ 60 \cdot \delta_{84}(n) + 48 \cdot \delta_{90}(n)
+ 60 \cdot \delta_{120}(n) + 48 \cdot \delta_{210}(n) \\
a_4(n)/n & = & (7n - 42)/12 \cdot \delta_6(n)
- 5/2 \cdot \delta_{12}(n) - 4 \cdot \delta_{18}(n)
+ 3 \cdot \delta_{24}(n) \\
& & \mbox{} + 6 \cdot \delta_{42}(n)
+ 34 \cdot \delta_{60}(n) - 6 \cdot \delta_{84}(n)
- 6 \cdot \delta_{120}(n) \\
a_5(n)/n & = & (n - 6)/4 \cdot \delta_6(n)
- 3/2 \cdot \delta_{12}(n) - 2 \cdot \delta_{24}(n)
+ 4 \cdot \delta_{42}(n) \\
& & \mbox{} + 6 \cdot \delta_{84}(n)
+ 6 \cdot \delta_{120}(n) \\
a_6(n)/n & = & 4 \cdot \delta_{30}(n)
- 4 \cdot \delta_{60}(n) \\
a_7(n)/n & = & \delta_{30}(n) + 4 \cdot \delta_{60}(n)
\end{eqnarray*}
hold for all the $n$ listed in Lemma~\ref{determined},
so the lemma implies that they hold for all $n \ge 3$.
These formulas imply the remarks in the introduction about the maximum
number of diagonals meeting at an interior point other than the center.
Finally
\begin{eqnarray*}
I(n) & = & \delta_2(n) + \sum_{k=2}^\infty a_k(n) \\
& = & \delta_2(n) + \sum_{k=2}^7 a_k(n), \\
\end{eqnarray*}
which gives the desired formula.
(The $\delta_2(n)$ in the expression for $I(n)$ is to account
for the center point when $n$ is even, which is the only point
not counted by the $a_k$.)
\end{pf*}
\section{The formula for regions}
\label{regions}
We now use the knowledge obtained in the proof of
Theorem~\ref{countintersections} about the number of interior points
through which exactly $k$ diagonals pass to calculate the number of
regions formed by the diagonals.
\begin{pf*}{Proof of Theorem~\ref{countregions}}
Consider the graph formed from the configuration of a regular $n$-gon
with its diagonals, in which the vertices are the vertices of the $n$-gon
together with the interior intersection points, and the edges are the
sides of the $n$-gon together with the segments that the diagonals cut
themselves into.
As usual, let $V$ denote the number of vertices of the graph, $E$ the
number of edges, and $F$ the number of regions formed, including the
region outside the $n$-gon.
We will employ Euler's Formula $V-E+F=2$.
Clearly $V=n+I(n)$.
We will count edges by counting their ends, which are $2E$ in number.
Each vertex has $n-1$ edge ends, the center (if $n$ is even) has $n$
edge ends, and any other interior point through which exactly $k$ diagonals
pass has $2k$ edge ends, so
$$2E = n(n-1) + n \delta_2(n) + \sum_{k=2}^\infty 2k a_k(n).$$
So the desired number of regions, not counting the region outside
the $n$-gon, is
\begin{eqnarray*}
F-1 & = & E-V+1 \\
& = & \left[ n(n-1)/2 + n \delta_2(n)/2 +
\sum_{k=2}^\infty k a_k(n) \right]
- \left[ n+I(n) \right] + 1.
\end{eqnarray*}
Substitution of the formulas derived in the proof of
Theorem~\ref{countintersections} for $a_k(n)$ and $I(n)$
yields the desired result.
\end{pf*}
\section*{Appendix: computations and tables}
\label{computations}
In Table~\ref{intersection_pts1}
we list $I(n), R(n), a_{2}(n),\ldots,a_{7}(n)$
for $n=4,5,\ldots,30$.
To determine the polynomials listed in
Theorem~\ref{countintersections}
more data was needed especially for $n \equiv 0 \bmod{6}$.
The largest
$n$ for which this was required was~420.
For speed and memory conservation,
we took advantage of the regular $n$-gon's rotational symmetry and focused
our attention on
only $2\pi/n$ radians of the $n$-gon. The data from this computation is
found in Table~\ref{intersection_pts2}. Although we only needed
to know the values at those $n$
listed in Lemma~\ref{determined} of Section~\ref{points},
we give a list for $n=6,12,\ldots,420$
so that the nice patterns can be seen.
The numbers in these tables were found by numerically computing (using a C
program and 64 bit precision) all possible $\left( {n}\atop{4} \right)$
intersections, and sorting them by their $x$ coordinate. We then
focused on runs of points with close $x$ coordinates, looking
for points with close $y$ coordinates.
Several checks were made to eliminate any fears (arising from round-off
errors) of distinct points being mistaken as close. First, the C program
sent data to Maple which
checked that the coordinates of close points agreed to at least
40 decimal places. Second, we verified for each $n$
that close points came in counts of the
form $\left( {k}\atop{2} \right)$ ($k$ diagonals
meeting at a point give rise to $\left( {k}\atop{2} \right)$
close points. Hence, any run whose length is not of this form
indicates a computational error).
A second program was then written and run on a second machine to make
the computations completely rigorous. It also found the intersection
points numerically, sorted them and looked for close points,
but, to be absolutely sure that a pair of close
points $p_1$ and $p_2$ were actually the same,
it checked that for the two pairs of diagonals $(l_1,l_2)$
and $(l_3,l_4)$ determining $p_1$ and $p_2$, respectively,
the triples $l_1,l_2,l_3$ and $l_1,l_2,l_4$ each divided the
circle into arcs of lengths consistent with Theorem~\ref{thesolutions}.
Since this test only involves comparing rational numbers,
it could be performed exactly.
A word should also be said concerning limiting the search to
$2\pi/n$ radians of the $n$-gon. Both programs looked
at slightly smaller slices of the $n$-gon to avoid problems
caused by points near the boundary. We further subdivided this
region into twenty smaller pieces to make the task of sorting the
intersection points manageable. More precisely,
we limited our search to points whose angle with the
origin fell between $[c_1+ 2\pi (m-1)/(20n)+\varepsilon,
c_1+ 2\pi m/(20n)-\varepsilon)$, $m=1,2,\ldots 20$, and also made sure
not to include the origin in the count.
Here $\varepsilon $ was chosen to be $.00000000001$
and $c_1$ was chosen to be $.00000123$ ($c_1=0$ would
have led to problems since there are many intersection points
with angle $0$ or $2\pi/n$). To make sure that
no intersection points were omitted, the number
of points found (counting multiplicity) was compared with
$({n \choose 4} - {{n/2} \choose 2} \delta_2)/n$.
\begin{table}
\begin{center}
\begin{tabular}{|l||l|l|l|l|l|l|l|l|}
$n$&$a_{2}(n)$&$a_{3}(n)$&$a_{4}(n)$&$a_{5}(n)$&$a_{6}(n)$&$a_{7}(n)$&
$I(n)$&$R(n)$ \\ \hline \hline
3 & & & & & & & 0 & 1\\
4 & & & & & & & 1 & 4\\
5 & 5 & & & & & & 5 & 11\\
6 & 12 & & & & & &13 & 24\\
7 & 35 & & & & & & 35 & 50\\
8 & 40 & 8 & & & & & 49 & 80\\
9 & 126 & & & & & & 126 & 154\\
10 & 140 & 20 & & & & & 161 & 220\\
11 & 330 & & & & & & 330 & 375\\
12 & 228 & 60 & 12 & & & & 301 & 444\\
13 & 715 & & & & & & 715 & 781\\
14 & 644 & 112 & & & & & 757 & 952\\
15 & 1365 & & & & & & 1365 & 1456\\
16 & 1168 & 208 & & & & & 1377 & 1696\\
17 & 2380 & & & & & & 2380 & 2500\\
18 & 1512 & 216 & 54 & 54 & & & 1837 & 2466\\
19 & 3876 & & & & & & 3876 & 4029\\
20 & 3360 & 480 & & & & & 3841 & 4500\\
21 & 5985 & & & & & & 5985 & 6175\\
22 & 5280 & 660 & & & & & 5941 & 6820\\
23 & 8855 & & & & & & 8855 & 9086\\
24 & 6144 & 864 & 264 & 24 & & & 7297 & 9024\\
25 & 12650 & & & & & & 12650 & 12926\\
26 & 11284 & 1196 & & & & & 12481 & 13988\\
27 & 17550 & & & & & & 17550 & 17875\\
28 & 15680 & 1568 & & & & & 17249 & 19180\\
29 & 23751 & & & & & & 23751 & 24129\\
30 & 13800 & 2250 & 420 & 180 & 120 & 30 & 16801& 21480\\
\hline
\end{tabular}
\end{center}
\caption{A listing of $I(n)$,$R(n)$ and $a_{2}(n),\ldots,a_{7}(n)$,
$n=3,4,\ldots,30$. Note that, when $n$ is even, $I(n)$ also counts
the point in the center.}
\label{intersection_pts1}
\end{table}
\begin{table}
\centerline{\small
\begin{tabular}{|l||l|l|l|l|l|l|l||l|l|l|l|l|l|l|l|}
$n$&$\frac{a_{2}(n)}{n}$&$\frac{a_{3}(n)}{n}$&$\frac{a_{4}(n)}{n}$
&$\frac{a_{5}(n)}{n}$&
$\frac{a_{6}(n)}{n}$&$\frac{a_{7}(n)}{n}$&$\frac{I(n)-1}{n}$&
$n$&$\frac{a_{2}(n)}{n}$&$\frac{a_{3}(n)}{n}$&$\frac{a_{4}(n)}{n}$
&$\frac{a_{5}(n)}{n}$&
$\frac{a_{6}(n)}{n}$&$\frac{a_{7}(n)}{n}$&$\frac{I(n)-1}{n}$
\\ \hline \hline
6 & 2 & & & & & & 2 &
216 & 392564 & 4848 & 119 & 49 & & & 397580 \\
12 & 19 & 5 & 1 & & & & 25 &
222 & 426836 & 5166 & 126 & 54 & & & 432182 \\
18 & 84 & 12 & 3 & 3 & & & 102 &
228 & 463303 & 5441 & 127 & 54 & & & 468925 \\
24 & 256 & 36 & 11 & 1 & & & 304 &
234 & 501762 & 5718 & 129 & 57 & & & 507666 \\
30 & 460 & 75 & 14 & 6 & 4 & 1 & 560 &
240 & 541612 & 6121 & 165 & 61 & & 5 & 547964 \\
36 & 1179 & 109 & 11 & 6 & & & 1305 &
246 & 584782 & 6340 & 140 & 60 & & & 591322 \\
42 & 1786 & 194 & 27 & 13 & & & 2020 &
252 & 629399 & 6693 & 137 & 70 & & & 636299 \\
48 & 3168 & 220 & 25 & 7 & & & 3420 &
258 & 676580 & 6972 & 147 & 63 & & & 683762 \\
54 & 4722 & 288 & 24 & 12 & & & 5046 &
264 & 725976 & 7276 & 151 & 61 & & & 733464 \\
60 & 6251 & 422 & 63 & 12 & & 5 & 6753 &
270 & 777420 & 7643 & 150 & 66 & 4 & 1 & 785284 \\
66 & 9172 & 460 & 35 & 15 & & & 9682 &
276 & 831575 & 7969 & 155 & 66 & & & 839765 \\
72 & 12428 & 504 & 35 & 13 & & & 12980 &
282 & 887986 & 8326 & 161 & 69 & & & 896542 \\
78 & 15920 & 642 & 42 & 18 & & & 16622 &
288 & 947132 & 8640 & 161 & 67 & & & 956000 \\
84 & 20007 & 805 & 43 & 28 & & & 20883 &
294 & 1008358 & 9056 & 174 & 76 & & & 1017664 \\
90 & 25230 & 863 & 45 & 21 & 4 & 1 & 26164 &
300 & 1072171 & 9462 & 203 & 72 & & 5 & 1081913 \\
96 & 31240 & 948 & 53 & 19 & & & 32260 &
306 & 1139436 & 9780 & 171 & 75 & & & 1149462 \\
102 & 37786 & 1096 & 56 & 24 & & & 38962 &
312 & 1208944 & 10164 & 179 & 73 & & & 1219360 \\
108 & 45447 & 1201 & 53 & 24 & & & 46725 &
318 & 1281100 & 10582 & 182 & 78 & & & 1291942 \\
114 & 53768 & 1368 & 63 & 27 & & & 55226 &
324 & 1356315 & 10957 & 179 & 78 & & & 1367529 \\
120 & 62652 & 1601 & 95 & 31 & & 5 & 64384 &
330 & 1434110 & 11375 & 189 & 81 & 4 & 1 & 1445760 \\
126 & 73676 & 1658 & 72 & 34 & & & 75440 &
336 & 1514816 & 11856 & 193 & 89 & & & 1526954 \\
132 & 85319 & 1825 & 71 & 30 & & & 87245 &
342 & 1598970 & 12216 & 192 & 84 & & & 1611462 \\
138 & 97990 & 2002 & 77 & 33 & & & 100102 &
348 & 1685843 & 12661 & 197 & 84 & & & 1698785 \\
144 & 112100 & 2136 & 77 & 31 & & & 114344 &
354 & 1775788 & 13108 & 203 & 87 & & & 1789186 \\
150 & 127070 & 2345 & 84 & 36 & 4 & 1 & 129540 &
360 & 1868312 & 13669 & 231 & 91 & & 5 & 1882308 \\
156 & 143635 & 2549 & 85 & 36 & & & 146305 &
366 & 1965272 & 14010 & 210 & 90 & & & 1979582 \\
162 & 161520 & 2736 & 87 & 39 & & & 164382 &
372 & 2064919 & 14465 & 211 & 90 & & & 2079685 \\
168 & 180504 & 3008 & 95 & 47 & & & 183654 &
378 & 2167754 & 14930 & 219 & 97 & & & 2183000 \\
174 & 201448 & 3178 & 98 & 42 & & & 204766 &
384 & 2274136 & 15396 & 221 & 91 & & & 2289844 \\
180 & 223251 & 3470 & 129 & 42 & & 5 & 226897 &
390 & 2383690 & 15885 & 224 & 96 & 4 & 1 & 2399900 \\
186 & 247562 & 3630 & 105 & 45 & & & 251342 &
396 & 2496999 & 16369 & 221 & 96 & & & 2513685 \\
192 & 273144 & 3844 & 109 & 43 & & & 277140 &
402 & 2613536 & 16896 & 231 & 99 & & & 2630762 \\
198 & 300294 & 4092 & 108 & 48 & & & 304542 &
408 & 2733888 & 17380 & 235 & 97 & & & 2751600 \\
204 & 329171 & 4357 & 113 & 48 & & & 333689 &
414 & 2857752 & 17898 & 234 & 102 & & & 2875986 \\
210 & 359556 & 4661 & 125 & 55 & 4 & 1 & 364402 &
420 & 2984383 & 18598 & 273 & 112 & & 5 & 3003371 \\
\hline
\end{tabular}
}
\vskip12pt
\caption{The number of intersection points for one piece
of the pie (i.e. $2\pi/n$ radians), $n=6,12,\ldots,420$.}
\label{intersection_pts2}
\end{table}
\section*{Acknowledgements}
We thank Joel Spencer and Noga Alon for helpful conversations.
Also we thank Jerry Alexanderson, Jeff Lagarias, Hendrik Lenstra,
and Gerry Myerson for pointing out to us many of the references below.
|
\section*{ABSTRACT}
This paper addresses an important problem in Example-Based Machine Translation
(EBMT), namely how to measure similarity between a sentence fragment and a set
of stored examples. A new method is proposed that measures similarity according
to both surface structure and content. A second contribution is the use of
clustering to make retrieval of the best matching example from the database
more efficient. Results on a large number of test cases from the CELEX database
are presented.
\section{INTRODUCTION}
EBMT is based on the idea of performing translation by imitating translation
examples of similar sentences [Nagao 84]. In this type of translation system, a
large amount of bi/multi-lingual translation examples has been stored in a
textual database and input expressions are rendered in the target language by
retrieving from the database that example which is most similar to the input.
There are three key issues which pertain to example-based translation:
\begin{itemize}
\item establishment of correspondence between units in a bi/multi-lingual text
at sentence, phrase or word level
\item a mechanism for retrieving from the database the unit that best matches
the input
\item exploit the retrieved translation example to produce the actual
translation of the input sentence
\end{itemize}
[Brown 91] and [Gale 91] have proposed methods for establishing correspondence
between sentences in bilingual corpora. [Brown 93], [Sadler 90] and [Kaji 92]
have tackled the problem of establishing correspondences between words and
phrases in bilingual texts.
The third key issue of EBMT, that is exploiting the retrieved translation
example, is usually dealt with by integrating into the system conventional MT
techniques [Kaji 92], [Sumita 91]. Simple modifications of the translation
proposal, such as word substitution, would also be possible, provided that
alignment of the translation archive at word level was available.
In establishing a mechanism for the best match retrieval, which is the topic of
this paper, the crucial tasks are: (i) determining whether the search is for
matches at sentence or sub-sentence level, that is determining the ``text
unit'', and (ii) the definition of the metric of similarity between two text
units.
As far as (i) is concerned, the obvious choice is to use as text unit the
sentence. This is because, not only are sentence boundaries unambiguous but
also translation proposals at sentence level is what a translator is usually
looking for. Sentences can, however, be quite long. And the longer they are,
the less possible it is that they will have an exact match in the translation
archive, and the less flexible the EBMT system will be.
On the other hand if the text unit is the sub-sentence we face one major
problem, that is the possibility that the resulting translation of the whole
sentence will be of low quality, due to boundary friction and incorrect
chunking. In practice, EBMT systems that operate at sub-sentence level involve
the dynamic derivation of the optimum length of segments of the input sentence
by analysing the available parallel corpora. This requires a procedure for
determining the best ``cover'' of an input text by
segments of sentences contained in the database [Nirenburg 93]. It is assumed
that the translation of the segments of the database that cover the input
sentence is known. What is needed, therefore, is a procedure for aligning
parallel texts at sub-sentence level [Kaji 92], [Sadler 90]. If sub-sentence
alignment is available, the approach is fully automated but is quite vulnerable
to the problem of low quality as mentioned above, as well as to ambiguity
problems when the produced segments are rather small.
Despite the fact that almost all running EBMT systems employ the sentence as
the text unit, it is believed that the potential of EBMT lies on the
exploitation of fragments of text smaller that sentences and the combination of
such fragments to produce the translation of whole sentences [Sato 90].
Automatic sub-sentential alignment is, however, a problem yet to be solved.
Turning to the definition of the metric of similarity, the requirement is
usually twofold. The similarity metric applied to two sentences (by sentence
from now on we will refer to both sentence and sub-sentence fragment) should
indicate how similar the compared sentences are, and perhaps the parts of the
two sentences that contributed to the similarity score. The latter could be
just a useful indication to the translator using the EBMT system, or a crucial
functional factor of the system as will be later
explained.
The similarity metrics reported in the literature can be characterised
depending on the text patterns they are applied on. So, the word-based metrics
compare individual words of the two sentences in terms of their morphological
paradigms, synonyms, hyperonyms, hyponyms, antonyms, pos tags... [Nirenburg 93]
or use a semantic distance d (0 $\leq$ d $\leq$ 1) which is determined by the
Most Specific Common Abstraction (MSCA) obtained from a thesaurus abstraction
hierarchy
[Sumita 91]. Then, a similarity metric is devised, which reflects the
similarity of two sentences, by combining the individual contributions towards
similarity stemming from word comparisons.
The word-based metrics are the most popular, but other approaches include
syntax-rule driven metrics [Sumita 88], character-based metrics [Sato 92] as
well as some hybrids [Furuse 92]. The character-based metric has been applied
to Japanese, taking advantage of certain characteristics of the Japanese. The
syntax-rule driven metrics try to capture similarity of two sentences at the
syntax level. This seems very promising, since similarity at the syntax level,
perhaps coupled by lexical similarity in a
hybrid configuration, would be the best the EBMT system could offer as a
translation proposal. The real time feasibility of such a system is, however,
questionable since it involves the complex task of syntactic analysis.
In section 2 a similarity metric is proposed and analysed. The statistical
system presented consists of two phases, the Learning and the decision making
or Recognition phase, which are described in section III. Finally, in section
IV the experiment configuration is discussed and the results evaluated.
\section{THE SIMILARITY METRIC}
To encode a sentence into a vector, we exploit information about the functional
words/phrases (fws) appearing in it, as well as about the lemmas and pos
(part-of-speech) tags of the words appearing between fws/phrases. Based on the
combination of fws/phrases data and pos tags, a simple view of the surface
syntactic structure of each sentence is obtained.
To identify the fws/phrases in a given corpus the following criteria are
applied:
\begin{itemize}
\item fws introduce a syntactically standard behaviour
\item most of the fws belong to closed classes
\item the semantic behaviour of fws is determined through their context
\item most of the fws determine phrase boundaries
\item fws have a relatively high frequency in the corpus
\end{itemize}
According to these criteria, prepositions, conjunctions, determiners, pronouns,
certain adverbials etc. are regarded as fws. Having identified the fws of the
corpus we distinguish groups of fws on the basis of their interchangeability in
certain phrase structures. The grouping caters, also, for the multiplicity of
usages of a certain word which has been identified as a fw, since a fw can be a
part of many different groups. In this way, fws can serve the retrieval
procedure with respect to the following
two levels of contribution towards the similarity score of two sentences :
\begin{itemize}
\item Identity of fws of retrieved example and input (I)
\item fws of retrieved example and input not identical but belonging to the
same group (G)
\end{itemize}
To obtain the lemmas and pos tags of the remaining words in a sentence, we use
a part-of-speech Tagger with no disambiguation module, since this would be time
consuming and not 100\% accurate. Instead, we introduce the concept of
ambiguity class (ac) and we represent each non-fw by its ac and the
corresponding lemma(s) (for example, the unambiguous word ``eat'' would be
represented by the ac which is the set {verb} and the lemma ``eat'') (in
English, for an ambiguous word, the corresponding lemmas will
usually be identical. But this is rarely true for Greek). Hence, the following
two levels of contribution to the similarity score stem from non-fws:
\begin{itemize}
\item overlapping of the sets of possible lemmas of the two words (L)
\item overlapping of the ambiguity classes of the two words (T)
\end{itemize}
Hence, each sentence of the source part of the translation archive is
represented by a pattern, which is expressed as an ordered series of the above
mentioned feature components.
A similarity metric is defined between two such vectors, and is used in both
the Learning and Recognition phases. Comparing a test vector against a
reference vector is, however, not straightforward, since there are generally
axis fluctuations between the vectors (not necessarily aligned vectors and of
most probably different length). To overcome these problems we use a two-level
Dynamic Programming (DP) technique [Sakoe 78], [Ney 84]. The first level treats
the matches at fw level, while the second is
reached only in case of a match in the first level, and is concerned with the
lemmas and tags of the words within fw boundaries. Both levels utilise the same
(DP) model which is next described.
We have already referred to the (I) and (G) contributions to the similarity
score due to fws. But this is not enough. We should also take into account
whether the fws appear in the same order in the two sentences, whether an extra
(or a few) fws intervene in one of the two sentences, whether certain fws are
missing ... To deal with these problems, we introduce a yet third contribution
to the similarity score, which is negative and is called penalty score (P). So,
as we are moving along a diagonal of the
xy-plane (corresponding to matched fws), whenever a fw is mismatched, it
produces a negative contribution to the score along a horizontal or vertical
direction. In figure 1 the allowable transitions in the xy-plane are shown:
\begin{figure}[htb]
\setlength{\unitlength}{4mm}
\begin{picture}(14,6)(0,0)
\put(7,0){\vector(1,1){4}}
\put(12,0){\vector(0,1){4}}
\put(7,5){\vector(1,0){4}}
\put(7,0){\line(1,1){5}}
\put(12,4){\line(0,1){1}}
\put(11,5){\line(1,0){1}}
\put(10,6){P}
\put(13,3){P}
\put(7.6,3){I (G)}
\end{picture}
\caption {The DP allowable transitions}
\end{figure}
Whenever a diagonal transition is investigated, the system calls the second
level DP-algorithm which produces a local additional score due to the potential
similarity of lemmas and tags of the words lying between the corresponding fws.
This score is calculated using exactly the same DP-algorithm as the one
treating fws (allowing additions, deletions,...), provided that we use (L), (T)
and (PT) (a penalty score attributed to a mismatch at the tag-level) in place
of (I), (G) and (P) respectively.
The outcome of the DP-algorithm is the similarity score between two vectors
which allows for different lengths of the two sentences, similarity of
different parts of the two sentences (last part of one with the first part of
the other) and finally variable number of additions and deletions. The score
produced, corresponds to two coherent parts of the two sentences under
comparison. Emphasis should be given to the variable number of additions and
deletions. The innovation of the penalty score (which is
in fact a negative score) provides the system with the flexibility to afford a
different number of additions or deletions depending on the accumulated
similarity score up to the point where these start. Moreover, the algorithm
determines, through a backtracking procedure, the relevant parts of the two
vectors that contributed to this score. This is essential for the sentence
segmentation described in the next section.
It should also be noted that the similarity score produced is based mainly on
the surface syntax of the two sentences (as this is indicated by the fws and
pos tags) and in the second place on the actual words of the two sentences.
This is quite reasonable, since the two sentences could have almost the same
words in the source language but no similarity at all in the source or target
language (due to different word order, as well as different word utilisation),
while if they are similar in terms of fws as
well as in terms of the pos tags of the words between fws, then the two
sentences would almost certainly be similar (irrelevant of a few differences in
the actual words) in the target language as well (which is the objective).
The DP-algorithm proposed seems to be tailored to the needs of the similarity
metric but there is yet a crucial set of parameters to be set, that is
A={I,G,P,L,T,PT}. The DP-algorithm is just the framework for the utilisation of
these parameters. The values of the parameters of A are set dynamically
depending on the lengths of the sentences under comparison. I, G, L, T are set
to values (I, G are normalised by the lengths of the sentences in fws, while L,
T are normalised by the lengths of the blocks of
words appearing between fws) which produce a 100\% similarity score when the
sentences are identical, while P, PT reflect the user's choise of penalising an
addition or deletion of a word (functional or not).
\section{LEARNING AND RECOGNITION PHASES}
In the Learning phase, the modified k-means clustering procedure [Wilpon 85] is
applied to the source part of the translation archive, aiming to produce
clusters of sentences, each represented by its centre only. The algorithm
produces the optimum segmentation of the corpus into clusters (based on the
similarity metric), and determines each cluster centre (which is just a
sentence of the corpus) by using the minmax criterion. The number of clusters
can be determined automatically by the process, subject
to some cluster quality constraint (for example, minimum intra-cluster
similarity), or alternatively can be determined externally based upon
memory-space restrictions and speed requirements.
Once the clustering procedure is terminated, a search is made, among the
sentences allocated to a cluster, to locate second best (but good enough)
matches to the sentences allocated to the remaining clusters. If such matches
are traced, the relevant sentences are segmented and then the updated corpus is
reclustered. After a number of iterations, convergence is obtained (no new
sentence segments are created) and the whole clustering procedure is
terminated.
Although the objective of a matching mechanism should be to identify in a
database the longest piece of text that best matches the input, the rationale
behind sentence segmentation is in this case self-evident. It is highly
probable that a sentence is allocated to a cluster center because of a good
match due to a part of it, while the remaining part has nothing to do with the
cluster to which it will be allocated. Hence, this part will remain hidden to
an input sentence applied to the system at the
recognition phase. On the other hand, it is also highly probable that a given
input sentence does not, as a whole, match a corpus sentence, but rather
different parts of it match with segments belonging to different sentences in
the corpus. Providing whole sentences as translation proposals, having a part
that matched with part of the input sentence, would perhaps puzzle the
translator instead of help him (her).
But sentence segmentation is not a straightforward matter. We can not just
segment a sentence at the limits of the part that led to the allocation of the
sentence to a specific cluster. This is because we need to know the translation
of this part as well. Hence, we should expand the limits of the match to cover
a ``translatable unit'' and then segment the sentence. Automatic sub-sentential
alignment (which would produce the ``translatable units''), however, is not yet
mature enough to produce high
fidelity results. Hence, one resorts to the use of semi-automatic methods (in
our application with the CELEX database, because of the certain format in which
the texts appear, a rough segmentation of the sentences is straightforward and
can therefore be automated).
If alignment at sub-sentential level is not available, the segmentation of the
sentences of the corpus is not possible (it is absolutely pointless). Then, the
degree of success of the Learning phase will depend on the length of the
sentences contained in the corpus. The longer these sentences tend to be, the
less successful the Learning phase. On the other hand, if alignment at
sub-sentential level is available, we could just apply the clustering procedure
to these segments. But then, we might end up
with an unnecessary large number of clusters and ``sentences''. This is
because, in a specific corpus quite a lot of these segments tend to appear
together. Hence, by clustering whole sentences and then segmenting only in case
of a good match with a part of a sentence allocated to a different cluster, we
can avoid the overgeneration of clusters and segments. When the iterative
clustering procedure is finally terminated, the sentences of the original
corpus will have been segmented to ``translatable
units'' in an optimum way, so that they are efficiently represented by a set of
sentences which are the cluster centres.
In the Recognition phase, the vector of the input sentence is extracted and
compared against the cluster centres. Once the favourite cluster(s) is
specified, the search space is limited to the sentences allocated to that
cluster only, and the same similarity metric is applied to produce the best
match available in the corpus. If the sentences in the translation archive have
been segmented, the problem is that, now, we do not know what the
``translatable units'' of the input sentence are (since we do not
know its target language equivalent). We only have potential ``translatable
unit'' markers. This is not really a restriction, however, since by setting a
high enough threshold for the match with a segment (translatable piece of text)
in the corpus, we can be sure that the part of the input sentence that
contributed to this good match, will also be translatable and we can,
therefore, segment this part. This process continues until the whole input
sentence has been ``covered'' by segments of the corpus.
\section{APPLICATION-EVALUATION}
The development of the matching method presented in this paper was part of the
research work conducted under the LRE I project TRANSLEARN. The project will
initially consider four languages: English, French, Greek and Portugese. The
application on which we are developing and testing the method is implemented on
the Greek-English language pair of records of the CELEX database, the
computerised documentation system on Community Law, which is available in all
Community languages. The matching mechanism is,
so far, implemented on the Greek part, providing English translation proposals
for Greek input sentences. The sentences contained in the CELEX database tend
to be quite long, but due to the certain format in which they appear
(corresponding to articles, regulations,...), we were able to provide the
Learning phase with some potential segmentation points of these sentences in
both languages of the pair (these segmentation points are in one-to-one
correspondence across languages, yielding the
``sub-sentence'' alignment).
In tagging the Greek part of the CELEX database we came across 31 different
ambiguity classes, which are utilised in the matching mechanism. The
identification and grouping of the Greek fws was mainly done with the help of
statistical tools applied to the CELEX database.
We tested the system on 8,000 sentences of the CELEX database. We are
presenting results on two versions. One of 80 clusters (which accounts for the
1\% of the number of the sentences of the corpus used) which resulted in 10,203
``sentences'' (sentences or segments) in 2 iterations, and one of 160 clusters
which resulted in 10,758 ``sentences'' in 2 iterations. To evaluate the system,
we asked five translators to assign each translation proposal of the system (in
our application these proposals sometimes
refer to segments of the input sentence) to one of four categories :
\noindent A : The proposal is the correct (or almost) translation\\
B : The proposal is very helpful in order to produce the translation\\
C : The proposal can help in order to produce the translation\\
D : The proposal is of no use to the translator\\
We used as test suite 200 sentences of the CELEX database which were not
included in the translation archive. The system proposed translations for 232
``sentences'' (segments or whole input sentences) in the former case and for
244 in the latter case. The results are tabulated in table 1 (these results
refer to the single best match located in the translation archive):
\begin{center}
Table 1
\begin{small}
\begin{tabular}{|l|l|l|} \hline
& {\bf 80 CLUSTERS} & {\bf 160 CLUSTERS} \\ \hline
{\bf A} & {\bf 220 (19\%)} & {\bf 244 (20\%)} \\ \hline
{\bf B} & {\bf 464 (40\%)} & {\bf 512 (42\%)} \\ \hline
{\bf C} & {\bf 209 (18\%)} & {\bf 245 (20\%)} \\ \hline
{\bf D} & {\bf 267 (23\%)} & {\bf 219 (18\%)} \\ \hline
& {\bf 1160} & {\bf 1220} \\ \hline
\end{tabular}
\end{small}
\end{center}
The table shows that in the case of 160 clusters, (1) at 62\% the system will
be very useful to the translator, and (2) some information can at least be
obtained from 82\% of the retrievals. In the case of 80 clusters the results do
not change significantly. Hence, as far as the similarity metric is concerned
the results seem quite promising (it should, however, be mentioned, that the
CELEX database is quite suitable for EBMT applications, due to its great degree
of repetitiveness).
On the other hand, the use of clustering of the corpus dramatically decreases
the response time of the system, compared to the alternative of searching
exhaustively through the corpus. Other methods for limiting the search space do
exist (for example, using full-text retrieval based on content words), but are
rather lossy, while clustering provides an effective means of locating the best
available match in the corpus (in terms of the similarity metric employed).
This can be seen in Table 2, where the
column ``MISSED'' indicates the percentage of the input ``sentences'' for
which the best match in the corpus was not located in the favourite cluster,
while the column ``MISSED BY'' indicates the average deviation of the located
best matches from the actual best matches in the corpus for these cases.
\begin{center}
Table 2
\begin{small}
\begin{tabular}{|l|l|l|} \hline
& {\bf MISSED} & {\bf MISSED BY} \\ \hline
{\bf 80 CLUSTERS} & {\bf 10\%} & {\bf 6.32\%} \\ \hline
{\bf 160 CLUSTERS} & {\bf 8.5\%} & {\bf 6.14\%} \\ \hline
\end{tabular}
\end{small}
\end{center}
In Table 1 as well as in Table 2 it can be seen that a quite important decrease
in the number of clusters affected the results only slightly. This small
deterioration in the performance of the system is due to ``hidden'' parts of
sentences allocated to clusters (parts that are not represented by the cluster
centres). Hence, the smaller the ``sentences'' contained in the database and
the more the clusters, the better the performance of the proposed system. The
number of clusters, however, should be
constrained for the search space to be effectively limited.
\section{REFERENCES}
[BROWN 91] Brown P. F. et al, (1991). ``Aligning Sentences in Parallel
Corpora''. {\em Proc. of the 29th Annual Meeting of the ACL}, pp 169-176.
[BROWN 93] Brown P. F. et al, (June 1993). ``The mathematics of Statistical
Machine Translation: Parameter Estimation''. {\em Computational Linguistics},
pp 263-311.
[FURUSE 92] Furuse O. and H. Iida, (1992). ``Cooperation between Transfer and
Analysis in Example-Based Framework''. {\em Proc. Coling}, pp 645-651.
[GALE 91] Gale W. A. and K. W. Church, (1991). ``A Program for Aligning
Sentences in Bilingual Corpora''. {\em Proc. of the 29th Annual Meeting of the
ACL.}, pp 177-184.
[KAJI 92] Kaji H., Y. Kida and Y. Morimoto, (1992). ``Learning Translation
Templates from Bilingual Text''. {\em Proc. Coling.}, pp 672-678.
[NAGAO 84] Nagao M., (1984). ``A framework of a mechanical translation between
Japanese and English by analogy principle''. {\em Artificial and Human
Intelligence}, ed. Elithorn A. and Banerji R., North-Holland, pp 173-180.
[NEY 84] Ney H., (1984). ``The use of a One-stage Dynamic Programming Algorithm
for Connected Word Recognition''. IEEE vol. ASSP-32, No 2.
[NIRENBURG 93] Nirenburg S. et al, (1993). ``Two Approaches to Matching in
Example-Based Machine Translation''. {\em Proc. of TMI-93, Kyoto, Japan}.
[SADLER 90] Sadler V. and R. Vendelmans, (1990). ``Pilot Implementation of a
Bilingual Knowledge Bank''. {\em Proc. of Coling}, pp 449-451.
[SAKOE 78] Sakoe H. and S. Chiba, (1978). ``Dynamic Programming Algorithm
Optimisation for Spoken Word Recognition''. {\em IEEE Trans. on ASSP, vol.
ASSP-26}.
[SATO 90] Sato S. and M. Nagao, (1990). ``Toward Memory-based Translation''.
{\em Proc. of Coling}, pp 247-252.
[SATO 92] Sato S., (1992). ``CTM: An Example-Based Translation Aid System''.
{\em Proc. of Coling}, pp 1259-1263.
[SUMITA 88] Sumita E. and Y. Tsutsumi, (1988). ``A Translation Aid System Using
Flexible Text Retrieval Based on Syntax-Matching''. {\em TRL Research Report},
Tokyo Research Laboratory, IBM.
[SUMITA 91] Sumita E. and H. Iida, (1991). ``Experiments and Prospects of
Example-based Machine Translation''. {\em Proc. of the 29th Annual Meeting of
the Association for Computational Linguistics}, pp 185-192.
[WILPON 85] Wilpon J. and L. Rabiner, (1985). ``A Modified k-Means Clustering
Algorithm for Use in Isolated Word Recognition''. {\em IEEE vol. ASSP-33}, pp.
587-594.
\end{document}
|
\section{Introduction}
The properties of matter at high temperature are interesting for a
number of experimental and cosmological applications. QCD at high
temperature and density may be relevant for heavy ion collisions,
while finite temperature phase transitions may play an important role
in the evolution of the universe. In gauge theories, an entirely analytic
perturbative study of the properties of high temperature matter is not
possible due to the so called infrared problem in the thermodynamics of
Yang-Mills fields~\cite{linde}.
A direct way to compute static equilibrium quantities at high
temperature would be to do lattice Monte Carlo
simulations in the 4d high
temperature theory. However, in many interesting cases the use of the
full 4d theory is difficult, if not impossible~\cite{FKRS2}.
These obstacles invoke a demand for a formalism which can solve in
a constructive way the problems mentioned. Since a finite
temperature equilibrium field theory is equivalent to a zero
temperature Euclidean field theory with compact 4th dimension, the
idea of the 4d $\rightarrow$ 3d dimensional reduction
is natural~\cite{G}--\cite{L}.
Dimensional reduction means that some properties of the
{\em equilibrium} high temperature plasma can be derived from a simpler
3d effective theory. The construction of the effective theory
is free of IR problems.
The 3d theory is purely bosonic,
and may then be studied by non-perturbative methods,
such as lattice MC simulations. In fact, the idea of dimensional
reduction has been around for quite
a long time~\cite{G,AP}. However, some concrete analytical results for the
construction of the 3d effective theory have appeared only recently. They
are relevant for the description of the high temperature electroweak phase
transition~\cite{FKRS2},\cite{FKRS1}--\cite{ml2} and high
temperature QCD~\cite{reisz1}--\cite{ay}.
The aim of the present paper is the formulation of the general rules
of dimensional reduction in a constructive way. Namely, we present
a set of 1-loop and 2-loop
Feynman diagrams with the results of their computation which can be
used for dimensional reduction in any gauge field theory.
As an example we construct the 3d effective theory corresponding to
the Standard Model of electroweak interactions.
New elements here in comparison with \cite{FKRS1}
are the inclusion of fermions, and the
direct relation of the parameters of the effective theory to the
physical parameters of the EW theory (the physical Z and
W boson, Higgs particle and top quark masses, the muon lifetime
and the temperature). We also discuss the
strategy for the derivation of the simplest possible effective theory
for typical extensions of the electroweak theory, like the models with
two Higgs doublets, and the Minimal Supersymmetric Standard Model.
The paper is organized as follows. In Section~2 we formulate the
general notion of dimensional reduction and analyse the expansion
parameters involved there. In Section~3 we present the building blocks for
the construction of the effective theory. Section~4 contains the dimensional
reduction of the Standard Model. In Section~5 we relate the parameters
in the $\overline{\mbox{\rm MS}}$ scheme to the physical parameters, thus completing the
relation of 3d couplings to temperature and observables.
Section~6 is a discussion. We argue there that
the effective theory of most of the extensions of Standard Model is
just the SU(2)$\times$U(1)+Higgs model.
\section{Dimensional reduction}
\label{DR}
The equilibrium properties of matter at high temperatures are related
to Matsubara Green's functions of different field operators.
By the concept of dimensional reduction we mean
that with some accuracy, all the 4d {\em static bosonic}
Green's functions in low energy domain (see below)
can be computed with the help of some effective 3d field theory.
Let us start with useful definitions.
\subsection{Superheavy, heavy and light modes}
\label{accuracy}
In order to define dimensional reduction,
consider a generic renormalizable field theory at high temperature
containing gauge $A_{\mu}$, scalar $\phi$ and fermionic fields
$\psi$,
\begin{equation}
L={1\over4} F_{\mu\nu}F_{\mu\nu} +
(D_{\mu}\phi)^\dagger(D_{\mu}\phi) + V(\phi) +
g_Y {\bar{\psi}}\phi\psi+\delta L.
\label{lagr}
\end{equation}
Here the group indices are suppressed, $\delta L$
contains the counterterms, and $V(\phi)$ is of the form\footnote{The
analysis of the case when cubic terms are present goes along
the same lines.}
\begin{equation}
V(\phi)=m_S^2\phi^\dagger\phi+\lambda(\phi^\dagger\phi)^2.
\end{equation}
For power counting
let us assume that $\lambda \sim g^2$ and $g_Y\sim g$, where $g$,
$\lambda$ and $g_Y$ are the gauge, scalar, and Yukawa couplings,
respectively. Write all 4d Matsubara fields in the form
\begin{equation}
\phi(x,\tau)=\sum_{n=-\infty}^{\infty}\phi_n(x)\exp(i
\omega^b_n \tau),
\end{equation}
\begin{equation}
\psi(x,\tau)=\sum_{n=-\infty}^{\infty}\psi_n(x)\exp(i
\omega^f_n \tau),
\end{equation}
where $\omega^b_n=2 n \pi T,~ \omega^f_n=(2n+1) \pi T $ are the 3d
tree masses for the bosonic ($\phi_n$) and fermionic ($\psi_n$) 3d
fields. Consider the 1-loop corrections
(for definiteness in the $\overline{\mbox{\rm MS}}$-scheme)
to the masses of the static
modes $\phi_0$ from the modes $\phi_{n\neq 0}$ and $\psi_n$.
In general, they have the form
\begin{equation}
m_i^2(T)=\gamma_i T^2 + m_i^2,~~ \gamma_i \sim g^2,
\label{mass}
\end{equation}
where $m_i^2$ is the zero-temperature mass of the scalar field
evaluated at some scale $\mu_T^m$ (see below, and~\cite{FKRS1}).
In general, $m^2(T)$ may be matrices, and in the
discussion below we mean the eigenvalues of those.
For the spatial components of the gauge fields $\gamma_i=0,~m_i^2=0$;
for the temporal components of the gauge
fields $\gamma_i\neq 0,~m_i^2=0$;
for the scalar fields $\gamma_i\neq 0,~ m_i^2 \neq 0$.
Now, let us divide the masses into different
categories depending on their magnitude at high temperature.
The 3d masses of all
fermionic modes and all bosonic modes with $n \neq 0$
are proportional to
$\pi T$, and we will call these modes superheavy.
The masses of the temporal components of the gauge fields~$A_0$
are proportional to $g T$, and these modes are called heavy.
The scalar fields can be separated in two different groups.
If $m_i^2$ is different from $-\gamma_i T^2$,
the scalar mass is proportional to $g T$, and the
field corresponding to this mass is ``heavy''.
In the contrary, one may be close to
a tree-level phase transition temperature so that
$\gamma_i T^2$ and $m_i^2$ cancel each other. Then
$m_i^2(T)\sim(g^2T)^2$, and
we call this field ``light''. We denote a generic light
scalar mass by $m_3^2$. All spatial components of
the gauge fields are ``light'' because for
them $\gamma_i=0$.
After these definitions we are ready to explain the
conjecture behind dimensional reduction. Two levels of dimensional
reduction are usually considered. On the first level the
effective theory is constructed for the light and heavy modes
(superheavy modes are ``integrated out''). The second level is the
theory for the light modes only.
In this paper we
require the 3d Lagrangian of the effective theory
to be super-renormalizable, so that
scalar self-interactions are at most quartic.
The super-renormalizable character introduces an
absolute upper bound on the accuracy of the description of
the 4d world by a 3d theory, to be discussed below.
\subsection{Two levels of dimensional reduction}
{\bf The theory for light and heavy modes.}
This theory is valid up
to momenta $k\ll T$, but $k$ may be as large as $g T$.
Consider a (super)renormalizable 3d gauge-Higgs theory with the
Lagrangian
\begin{equation}
L={1\over4} F_{ij}F_{ij} +
(D_{i}\phi)^\dagger(D_{i}\phi) + V_3(\phi,A_0)+ \fr12 (D_{i}A_0)^2
+\delta L,
\label{3dheavy}
\end{equation}
where $V_3(\phi,A_0)$ is of the form
\begin{equation}
V_3(\phi,A_0)=m^2\phi^\dagger\phi+\lambda_3(\phi^\dagger\phi)^2+
h_3\phi^\dagger\phi A_0^2+\frac{1}{2}m_D^2A_0^2+
\frac{1}{4}\lambda_A A_0^4.
\end{equation}
The gauge couplings $g_{3}$ have the dimension
GeV$^{1/2}$ and the scalar couplings $\lambda_{3}$
the dimension GeV.
To leading order, the parameter $m_D\sim gT$
is nothing but the Debye mass.
Consider bosonic static
$n$-point one-particle-irreducible Matsubara Green's functions
$G^{(4)}_{n}(\vec{k}_i)$ for the light and heavy fields in the full
4d theory, multiplied by factor $T^{n/2-1}$ to have the
dimension GeV$^{3-n/2}$, and depending on external
3-momenta $\vec{k}_i$. The statement of dimensional reduction is that
there is a mapping of the temperature
and the 4d coupling constants of the underlying theory
to the 3d theory such that the 3d theory gives
the same light and heavy Green's functions as the full 4d theory for
$k\le gT$ up to terms of order $O(g^4)$,
\begin{equation}
\frac{\Delta G}{G} \sim O(g^4).
\label{accurh}
\end{equation}
Fourth order in $g$ appears from a powercounting estimate
of the contributions of the neglected 6-dimensional operators to
typical Green's functions. For example, the operator $g^6 \phi^2 A_0^4/T^2$
contributes to the 2-point scalar correlator at order $g^6 m_D^2\sim g^8T^2$.
Since the order of magnitude of $m_3^2$ is $g^4 T^2$, the relative
error is $O(g^4)$. The same estimate arises by comparing
the contribution of the neglected operator $g^2 (k^4/T^2)\phi^2$
to the tree-level term $k^2$ at momenta $k\sim gT$.
To reach the accuracy goal~(\ref{accurh}),
the parameters of the 3d theory should be known
with relative uncertainty~$O(g^4)$, which means
1-loop accuracy~[$T(g^2+g^4)$] for the coupling constants,
2-loop accuracy~[$T^2(g^2+g^4)$] for the heavy
masses, and 3-loop accuracy~[$m^2+T^2(g^2+g^4+g^6)$] for the
light scalar masses.
Some comments are now in order.\\
(i) The problem of constructing an
effective 3d theory giving an accuracy better than $O(g^4)$ for {\em all}
Green's functions is far from being trivial (if possible at all).
It is clear, though, that if the theory exists, it must contain 6-dimensional
operators, and the 4d-3d mapping for the light scalar modes must
be done beyond 3-loop level.\\
(ii) Often dimensional reduction is done on the tree-level for
the couplings and 1-loop level
for the masses, i.e., at order $g^2$. This 3d theory
reproduces the 1-loop resummed effective
potential for the Higgs field~\cite{DHLLL,Ca,BFHW}.
However, the relative uncertainty in the mass squared
of the light scalar field is
$\Delta m_3^2/m_3^2\sim O(1)$, since the
tree-level mass term is compensated for by the 1-loop
thermal correction near the phase transition.
Hence ${\Delta G}/{G}\sim O(1)$,
and from the point of view of calculating general
correlators, the theory is useless.
To obtain the minimal useful accuracy $O(g^2)$,
one should go to the 2-loop order $g^4$
in the scalar mass parameter. A more complete $g^4$ calculation,
including 1-loop dimensional reduction [$T(g^2+g^4)$] for
the couplings coupled to the scalar fields, 1-loop
dimensional reduction [$T^2g^2$] for the heavy masses,
and 2-loop dimensional
reduction [$m^2+T^2(g^2+g^4)$] for the scalar mass,
is needed~\cite{FKRS1} to
reproduce the resummed 2-loop effective potential
for the Higgs field~\cite{AE,FH}.
The accuracy $g^4$ corresponds to 1-loop accuracy
in vacuum renormalization, and we will work
with this accuracy throughout this paper. In a weakly
coupled theory, the relative error $O(g^2)\sim g^2/16\pi^2$
of the $g^4$-calculation
is numerically very small. In the Standard Model, the
largest contributions arise from the top quark.\\
(iii) For some quantities, such as the critical temperature and the
observables in the broken phase, the $g^4$ computation
described in (ii) gives a relative error of order $O(g^4)$.
Consider now the second level of dimensional reduction.
{\bf The theory for light modes only.} This theory is valid up to
momenta $k\ll gT$, but $k$ may be as large as $g^2 T$. The Lagrangian
for this theory is just
\begin{equation}
L={1\over4} F_{ij}F_{ij} +
(D_{i}\phi)^\dagger(D_{i}\phi) + V_3(\phi),
\label{3dlight}
\end{equation}
where $V_3(\phi)$ is of the form
\begin{equation}
V_3(\phi)=\bar{m}_3^2\phi^\dagger\phi+\bar{\lambda}_3(\phi^\dagger\phi)^2.
\label{3dlV}
\end{equation}
Only light scalar fields are present.
The effective field theory
can provide the accuracy
\begin{equation}
\frac{\Delta G}{G} \sim O(g^3).
\label{accurl}
\end{equation}
This estimate arises as follows:
there are neglected 6-dimensional operators
of the form ${g^6 T\phi^6}/{m_{\rm heavy}^3}\sim
{g^3\phi^6}/{T^2}$, contributing to the
two-point scalar correlator at order $g^3m_3^2\sim g^7 T^2$.
This should be compared with $m_3^2\sim g^4T^2$.
Note that in contrast to the integration over the superheavy scale,
odd powers of coupling constants appear, since $m_{\rm heavy} \sim gT$.
To reach the accuracy (\ref{accurl}) one must know
$\bar{\lambda}_3$ in eq.~\nr{3dlV} including corrections
of order $g^4T$ and $\bar{m}_3^2$ including corrections of
order $g^6T^2$.
Comments analogous to the above are applicable to the second level of
dimensional reduction:\\
(ii) To go beyond the accuracy $O(g^3)$, the 6-dimensional operators
must be included in the Lagrangian and light scalar masses must
be computed at least with accuracy $g^7T^2$.\\
(ii) In practice, it is convenient to do the integration over
the heavy scale to the same order in the loop expansion as
the integration over the superheavy scale. This means 1-loop
level [$T(g^2+g^3)$] for the couplings and
2-loop level [$m_3^2+T^2(g^3+g^4)$] for
the scalar mass squared~$\bar{m}_3^2$.
The relative error in the couplings is then $O(g^2)$.
In~$\bar{m}_3^2$, the relative error is $O(g)$, which
is also the relative error in the Green's functions.
Numerically, $O(g)\sim g/4 \pi$ is small in a weakly coupled
theory. Note that since the theory in eq.~\nr{3dheavy} is purely bosonic,
there are no large fermionic corrections.\\
(iii) The procedure described in (ii) provides $O(g^3)$ accuracy in the
critical temperature and the broken phase observables.
Concrete numerical estimates of the accuracy of the effective
field theory depend on the observable and on the details of the
model. Some estimates for the electroweak theory were presented in
\cite{FKRS2,FKRS1,JKP}, and we add some more in Section 5.4.
\subsection{Dimensional reduction by matching}
The definition of dimensional reduction described above provides a
method of mapping the 4d theory on the 3d one. One just writes down
the most general 3d super-renormalizable Lagrangian for the heavy and
light modes, and defines its parameters by matching to a specified
accuracy the 2-, 3-,
and 4-point Green's functions in the 3d effective theory and in the
underlying 4d fundamental theory. The Green's functions to be matched
correspond to those appearing in the 3d Lagrangian. For the
2-point functions one needs the momentum dependent part,
but the 4-point functions may be taken at vanishing external momenta.
Due to gauge invariance, the 3-point
functions are not needed at all.
The scalar Green's functions with vanishing
momenta are most conveniently generated from
an effective potential.
Consider in some more detail the renormalized
2-point function for the light scalar field. In the full 4d
theory, it is of the form
\begin{equation}
k^2+m_S^2+\Pi(k^2)=k^2+m_S^2+\Pi_3(k^2)+\overline{\Pi}(k^2),\label{4dp}
\end{equation}
and we want to match it to the corresponding function in the 3d theory:
\begin{equation}
k^2+m_3^2+\Pi_3(k^2). \label{3dres}
\end{equation}
Here $\Pi_3(k^2)$ is the contribution of the light and heavy modes only,
and $\overline{\Pi}(k^2)$ represents all other contributions
(corresponding 2-loop graphs contain at least 2 superheavy
lines)\footnote{To be precise, one must use
resummation to produce the correct
$\Pi_3(k^2)$ in eq.~\nr{4dp};
however, this is not relevant for the present argument.}.
Since there are no IR-problems related to
the integration over the superheavy modes,
$\overline{\Pi}(k^2)$ is analytic in the external momentum~$k^2$,
and can for $k\ll T$ be expanded as
\begin{equation}
\overline{\Pi}(k^2)=\overline{\Pi}(0)+\overline{\Pi}'(0)k^2+
O(g^2\frac{k^4}{T^2}).\label{pik2}
\end{equation}
Here the $\Pi$'s are of order $g^2$ and if we restrict to $k\le gT$,
the higher-order contributions $O(g^2{k^4}/{T^2})$
are at most of order $g^6T^2$ and can be
neglected.
Assuming that $\overline{\Pi}(0)$ has been calculated
to 2-loop accuracy [$T^2(g^2+g^4)$] and
$\overline{\Pi}'(0)$ to 1-loop accuracy $g^2$, one can
rewrite the right-hand-side of eq.~\nr{4dp} as
\begin{equation}
[1+\overline{\Pi}'(0)]\Bigl\{
k^2+[m_S^2+\overline{\Pi}(0)][1-\overline{\Pi}'(0)]+\Pi_3(k^2)
\Bigr\},\label{4dres}
\end{equation}
where $\Pi_3(k^2)$ is of order $g^2T m_D\sim g^3T^2$ and
only terms up to order $g^4T^2$ are kept.
The matching of eqs.~\nr{3dres}
and \nr{4dres} can now be carried out by relating the
normalizations of the fields in 3d and 4d through
\begin{equation}
\phi_\rmi{3d}^2={1\over T}[1+\overline{\Pi}'(0)]\phi_{4d}^2,
\end{equation}
and by relating the masses as
\begin{equation}
m_3^2=[m_S^2+\overline{\Pi}(0)][1-\overline{\Pi}'(0)],
\end{equation}
which is the order $g^4$ result for $m_3^2$.
The other coupling constants can be fixed
similarly, using the appropriate correlators and
taking always into account the different normalizations
of the fields in 4d and 3d. The 3d theory relevant
for the Standard Model is constructed in this way
in Sec.~\ref{DRinSM}.
The general structure of the
relationships of the 4d and 3d parameters
is determined by
the super-renormalizable character of the 3d theory.
The 4d couplings and masses
are functions of the 4d $\overline{\mbox{\rm MS}}$ parameter~$\mu_4$, but the 3d scalar
and gauge coupling constants are renormalization group (RG) invariant,
since the 3d theory contains only mass divergences. For example, on
the 1-loop level the relationships of the 4d and 3d coupling
constants must have the form
\begin{equation}
g_3^2 = T[g^2(\mu_4) - \beta_{g^2}\log(\mu_4/c_{g^2} T)],\quad
\lambda_3= T[\lambda(\mu_4) - \beta_\lambda\log(\mu_4/c_{\lambda}T)],
\end{equation}
where $c_{g^2}$ and $c_{\lambda}$ are
definite fixed functions of physical parameters computable in
perturbation theory (see below),
and the $\beta$'s are the corresponding $\beta$-functions.
The scalar masses in the effective 3d theory, on the other hand,
are not RG-invariant,
but require ultraviolet renormalization on the 2-loop level.
Just dimensionally, the renormalized mass
parameters are of the form
\begin{equation}
m_3^2(\mu_3)= {1\over16\pi^2}f_{2m}
\log\frac{\Lambda_m}{\mu_3},
\label{3dmass}
\end{equation}
where $f_{2m}\sim g_3^4$. For clarity, let us point out
that $\mu_3$ in eq.~\nr{3dmass} is independent of the $\mu_4$
of the 4d theory,
since the bare mass parameter produced by the
dimensional reduction step is RG-invariant.
In Sec.~\ref{blocks} we present a
set of rules, together with a
computation of the necessary Feynman diagrams,
allowing one to define the mapping
of 4d on 3d (at 1-loop level for
coupling constants and 2-loop level for masses)
for an arbitrary gauge theory.
An important comment is now in order. The matching
procedure of dimensional reduction described above
is {\em different} from the initial \cite{G} method
of dimensional reduction which is defined as the sequence of
the following steps: \\
(i) Define a 3d bosonic effective action as
\begin{equation}
\exp(-S_\rmi{eff})=\int D\psi D\phi_{n \neq 0}\exp(-S),
\label{io}
\end{equation}
where integration over all superheavy modes is performed.\\
(ii)Make a perturbative computation of $S_\rmi{eff}$ and represent it
in
the form
\begin{equation}
S_\rmi{eff}= c V T^3+\int d^3x L_{\rmi{eff}}(T) + \sum_n
\frac{O_n}{T^n},
\end{equation}
where $L_{\rmi{eff}}(T)$ is a renormalizable 3d effective bosonic
Lagrangian with temperature-dependent constants, $O_n$ are operators
of dimensionality $n$, suppressed by powers of temperature, $c$ is
a number related to the number of degrees of freedom of the theory
and $V$ is the volume of the system.\\
(iii) Drop all the terms $O_n$. The effective action contains then
light and heavy fields. The final step is the integration over the
heavy modes in a way described in (i).
\vspace*{0.5cm}
The difficulties with the procedure described above,
arising at 2-loop level, have been pointed
out, e.g.,\ in~\cite{J,Mack}.
The problems are due to steps (ii) and (iii), since
step (i) produces non-local operators which cannot
be expanded in powers of $p^2/T^2$. In terms of
graphs, in the procedure of eq.~\nr{io} the internal
lines of the Feynman diagrams are always superheavy (or heavy). For
example, the only scalar diagram contributing to the scalar mass
renormalization on the 2-loop level is shown in Fig.~\ref{dr2l}.a.
In the Green's function approach the extra graph in
Fig.~\ref{dr2l}.b, containing two superheavy and one light
internal lines appears. As is pointed out in~\cite{FKRS1} this
diagram does not vanish in the high temperature limit,
and therefore, gives a contribution to the 3d mass.
Physically, the reason is that light fields
can have high momenta $p\sim T$ when they interact with
the superheavy fields. The need to include light fields
in the internal lines of many-loop graphs in order to
establish a useful local effective field theory,
is also well known in the context of large-mass expansion in
zero-temperature field theory (see, e.g., \cite{Co}).
When we speak of ``integrating over'' the superheavy or heavy
scale below, we always mean the matching procedure
for the Green's functions described in this Section.
\section{Building blocks for dimensional reduction}
\label{blocks}
In this Section, we give results for the typical diagrams
appearing in the construction of the effective 3d theory.
We account here for the momentum integrations and spin
contractions; the isospin contractions,
combinatorial factors, and coupling constants
relevant for the Standard Model
are added in Sec.~\ref{DRinSM}.
We work in Landau gauge, where the vector propagators
are transversal. The wave function normalization factors
relating the 4d and 3d fields depend
on the gauge condition~\cite{L}, but the final parameters
of 3d theory are gauge-independent at least to the order
in which we are working~\cite{FKRS1,JP}.
Landau-gauge is a convenient choice
since it reduces the number of diagrams considerably:
an external scalar leg with vanishing momentum cannot
directly couple to a vector field, since the vertex is
proportional to the loop momentum,
and hence gives zero when contracted with
the transversal vector propagator.
We will work throughout in Euclidian space.
The conventions for the Euclidian $\gamma$-matrices $\gamma_\mu$
in terms of the Minkowskian matrices $\gamma^\mu$ are
that $\gamma_0=\gamma^0$, $\gamma_i=-i\gamma^i$. The main
properties are $\gamma_\mu^\dagger=\gamma_\mu$,
$\{\gamma_\mu,\gamma_\nu\}=2\delta_{\mu\nu}$,
$\mathop{\rm Tr}\gamma_\mu=0$, $\mathop{\rm Tr} 1=4$.
Due to the relations $t=-i\tau$ and $A_0^M=iA_0^E$
between Minkowskian and Euclidian variables,
the covariant derivative is
$i\gamma^\mu D_\mu^M= -\gamma_\mu D_\mu^E$.
The matrix $\gamma_5$ satisfies
\begin{eqnarray}
& & \{\gamma_\mu,\gamma_5\} = 0,\quad\gamma_5^2=1,\quad
\nonumber \\
& & \mathop{\rm Tr}\gamma_5 =
\mathop{\rm Tr}\gamma_5\gamma_\mu\gamma_\nu=0,\quad
\mathop{\rm Tr}
\gamma_5\gamma_\mu\gamma_\nu\gamma_\sigma\gamma_\rho
\propto \epsilon_{\mu\nu\rho\sigma}.
\end{eqnarray}
We define $a_{R,L}=(1\pm\gamma_5)/2$.
The general form of the theory is the following.
There are scalars $\phi$, vector fields~$A^a_\mu$,
ghosts $\eta^a$, and fermions $\psi$. In the symmetric phase, only
the scalar fields have a mass parameter; any mass
parameters are inessential to dimensional reduction, though,
since we assume $m\sim gT$ so that masses contribute
at higher order . The propagators are
\begin{eqnarray}
\langle\phi(-p)\phi(p)\rangle & = & \frac{1}{p^2+m_S^2},\quad
\langle A^a_\mu(-p)A^b_\nu(p)\rangle=
\delta^{ab}\frac{\delta_{\mu\nu}-\frac{p_\mu p_\nu}
{p^2}}{p^2}, \nonumber\\
\langle \bar{\eta}^a(p) \eta^b(p)\rangle & = &
-\frac{\delta^{ab}}{p^2},\quad\quad\quad
\langle \bar{\psi}_\alpha(p)\psi_\beta(p)\rangle=
\frac{i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}_{\beta\alpha}}{p^2}. \label{symprop}
\end{eqnarray}
Defining
\begin{eqnarray}
F_{\mu\nu\rho}(p,q,r) & = &
\delta_{\mu\rho}(p_\nu-r_\nu)+
\delta_{\rho\nu}(r_\mu-q_\mu)+
\delta_{\nu\mu}(q_\rho-p_\rho), \\
G^{abcd}_{\mu\nu\sigma\!\rho} & = &
f^{abe}f^{cde}(\delta_{\mu\sigma}\delta_{\nu\rho}-
\delta_{\mu\rho}\delta_{\nu\sigma})+
(b\leftrightarrow c,\nu\leftrightarrow\sigma)+
(b\leftrightarrow d,\nu\leftrightarrow\rho),\nonumber
\end{eqnarray}
where $f^{abc}$ is antisymmetric,
the theory has the following types of vertices.
The self-interactions of vector fields are due to
vertices of the form
\begin{eqnarray}
\mbox{ } & &
igf^{abc}F_{\mu\nu\rho}(p,q,r)A^a_\mu(p)A^b_\nu(q)A^c_\rho(r),\quad
igf^{abc}p_\mu\bar{\eta}^a(p)A^b_\mu(q)\eta^c(r),\nonumber \\
& & g^2G^{abcd}_{\mu\nu\sigma\!\rho}A^a_\mu A^b_\nu A^c_\sigma A^d_\rho.
\label{gaugevs}
\end{eqnarray}
In the actual calculation one only needs the expression
\begin{equation}
G^{\alpha\alpha cd}_{\mu\nu\sigma\!\rho} =
f^{\alpha ce}f^{\alpha de}(2\delta_{\mu\nu}\delta_{\sigma\!\rho}-
\delta_{\mu\sigma}\delta_{\nu\rho}-
\delta_{\mu\rho}\delta_{\nu\sigma}),
\end{equation}
where $\alpha$ is not summed over,
so that the isospin part separates
for the quartic vertex as it does for the cubic one.
Fermions interact through vertices of the type
\begin{equation}
ig\bar{\psi}\gamma_\mu A_\mu a_L\psi,\quad
ig\bar{\psi}\gamma_\mu A_\mu \psi,\quad
g_Y \bar{\psi}\phi\psi,\quad
g_Y \bar{\psi}\gamma_5\phi\psi,\label{fv}
\end{equation}
and the scalar vertices are of the form
\begin{equation}
\lambda \phi^4,\quad
ig(p_\mu-r_\mu)\phi(p)A_\mu(q)\phi(r),\quad
g^2\phi\phi A_\mu A_\mu.\label{sv}
\end{equation}
In the above formulas, momentum conservation is implied.
The isospin indices are suppressed in eqs.~\nr{fv} and~\nr{sv}.
It turns out that for the calculations in this paper
it is sufficient to treat explicitly only the first and third
vertex in eq.~\nr{fv}, since the other two give
results differing only by trivial numerical coefficients.
In addition to the renormalized vertices,
one needs counterterms. The wave function
counterterms are denoted by
$\delta Z_S=Z_S-1$ (and similarly for the other fields),
where
\begin{equation}
\phi_B=Z_S^{1/2}\phi,\quad
A_B=Z_V^{1/2}A,\quad
\psi_{L,B}=(Z_F^L)^{1/2}\psi_L,\quad
\psi_{R,B}=(Z_F^R)^{1/2}\psi_R
\end{equation}
and $\psi_{L(R)}=a_{L(R)}\psi$. The only mass counterterm
in the symmetric phase is $\delta m_S^2$.
In the broken phase, the shift in the scalar field
generates mass counterterms for vectors and fermions, as well.
The coupling constant counterterms are denoted by $\delta g^2$,
$\delta \lambda$ and $\delta g_Y$, and are defined by
\begin{eqnarray}
& &
g_B^2\phi_B^2 A_B^2=(g^2+\delta g^2)\phi^2A^2,\quad
\lambda_B\phi_B^4=(\lambda+\delta \lambda)\phi^4,\nonumber \\
& & g_{Y,B}\bar{\psi}_B\phi_B\psi_B=(g_Y+\delta g_Y)
\bar{\psi}\phi\psi.
\end{eqnarray}
\subsection{Integration over the superheavy scale}
In this Section
we construct a local 3d effective field
theory which contains the bosonic
$n=0$ Matsubara modes only, and produces
the same static Green's functions as the full 4d theory
with the required accuracy.
As explained in Sec.~\ref{DR},
the recipe is to first identify the
general structure of the effective theory, and then to
compare static correlators calculated from the 3d
and 4d theories. The structure of the effective theory
differs from the tree-level action for $n=0$ modes
in the 4d theory in that the absence of Lorentz symmetry
allows the temporal components of the gauge fields to
develop mass terms and quartic self-interactions.
At 1-loop level, the construction of the 3d theory
proceeds simply by calculating the effect of
fermions and $n\neq 0$ bosons to two-, \mbox{three-,} and
four-point correlators of the static modes.
At 2-loop level, there can be $n=0$ modes in the
loops, as well, and hence one must carefully compare the
correlators in the two theories.
In Sec.~\ref{3dwf}
we calculate how the 3d fields are related to the 4d fields,
in Sec.~\ref{3dgs} we compute the effective couplings of the gauge sector,
in Sec.~\ref{fundam} we address the fundamental scalar sector, and in
Sec.~\ref{adjoint} we study the adjoint scalar sector, which is
composed of the temporal
components of the gauge fields.
\subsubsection{Notation and basic integrals}
To give results for the diagrams appearing in the integration
over the superheavy fields, we use the following notation:
\begin{eqnarray}
\Tint{p} & = & T\sum_n\int\frac{d^dp}{(2\pi)^d},\quad
\Tint{p}' = T\sum_{n\neq 0}\int\frac{d^dp}{(2\pi)^d},\quad
d=3-2\epsilon,\nonumber \\
p_b & = & (\omega_n^b,\vec{p})
,\quad p_f=(\omega_n^f,\vec{p}),\quad
\omega^b_n=2 n \pi T,\quad \omega^f_n=(2n+1) \pi T,\quad
k \equiv (0,\vec{k}), \nonumber \\
\imath_\epsilon & = &\ln\frac{\mu^2}{T^2}+2 \gamma_E-2\ln 2-
2\frac{\zeta'(2)}{\zeta(2)},\quad
c=\frac{1}{2}\biggl[\ln \frac{8\pi}{9}+\frac{\zeta'(2)}{\zeta(2)}-
2\gamma_E]\approx -0.348725, \nonumber \\
{c_B} & = & \ln (4\pi)-\gamma_E \approx
1.953808,\quad c_F=c_B-2\ln 2\approx
0.567514, \nonumber \\
L_b & = & \ln\frac{\mu^2}{T^2}-2 c_B,\quad
L_f=\ln\frac{\mu^2}{T^2}-2 c_F,\quad
\frac{1}{\epsilon_b}=\frac{1}{\epsilon}+L_b,\quad
\frac{1}{\epsilon_f}=\frac{1}{\epsilon}+L_f.\label{nabi}
\end{eqnarray}
The theory is regularized in the $\overline{\rm MS}$-scheme,
$\mu$ is the corresponding scale parameter.
The basic integrals appearing in 1-loop
integration over the superheavy modes are the following.
The fermionic and bosonic tadpole integrals, to the
accuracy they are needed, are~\cite{AE}
\begin{eqnarray}
I_b'(m) & = & \Tint{p_b}' \frac{1}{p^2+m^2}=\mu^{-2\epsilon}
\biggl[\frac{T^2}{12}(1+\epsilon\imath_\epsilon)-
\frac{m^2}{16\pi^2}\biggl(\frac{1}{\epsilon}+L_b\biggr)
\biggr], \label{Ib} \\
I_f(m) & = & \Tint{p_f} \frac{1}{p^2+m^2}=\mu^{-2\epsilon}
\biggl[-\frac{T^2}{24}\Bigl[1+\epsilon
(\imath_\epsilon-2\ln 2)\Bigr]-
\frac{m^2}{16\pi^2}\biggl(\frac{1}{\epsilon}+L_f\biggr) \label{If}
\biggr].
\end{eqnarray}
Taking derivatives with respect to mass squared and temperature
in eqs.~\nr{Ib}, \nr{If}, one can derive other integrals.
In the end one can put the masses in the propagators
to zero, since the integrals
over superheavy modes are analytic in the mass parameters, and
hence the effect of higher orders is suppressed by $m^2/T^2$.
The dependence on external momenta is likewise
analytic, and can be expanded in $k^2/T^2$.
Since all the parameters of the
effective theory are at most of order $gT$,
higher order contributions in $k^2/T^2$
can only produce contributions suppressed
by coupling constants.
The masses will play a role only in Sec.~\ref{fundam},
where we calculate integrals over superheavy modes
not directly, but by using the effective potential;
the needed integrals are given there.
The required massless integrals are
\begin{eqnarray}
B_b'& \equiv & \Tint{p_b}'\frac{1}{(p^2)^2}=
\frac{1}{16\pi^2}\frac{1}{\epsilon_b}, \nonumber \\
B_b'(k) & = & \Tint{p_b}'\frac{1}{p^2(p+k)^2} =
\frac{1}{16\pi^2}\frac{1}{\epsilon_b}
\biggl[1+{\cal O}\Bigl(\frac{k^2}{T^2}\Bigr)\biggr], \nonumber \\
J^b_{\alpha\beta}
& \equiv &
\Tint{p_b}'\frac{p_\alpha p_\beta}{p^2(p+k)^2}
-\Tint{p_b}'\frac{p_\alpha p_\beta}{(p^2)^2}, \nonumber \\
J^b_{00} & = & - \frac{k^2}{16\pi^2}
\biggl(\frac{1}{12\epsilon_b}+\frac{1}{6}\biggr), \nonumber \\
J^b_{ij} & = & - \biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{k^2}{16\pi^2}
\biggl(\frac{1}{12\epsilon_b}\biggr)+
\frac{k_ik_j}{16\pi^2}
\biggl(\frac{1}{4\epsilon_b}\biggr),
\label{integrals} \\
K^b_{\alpha\beta}
& \equiv &
\Tint{p_b}'\frac{p_\alpha p_\beta}{(p^2)^2(p+k)^2}, \nonumber \\
K^b_{00} & = & \frac{1}{16\pi^2}
\biggl(\frac{1}{4\epsilon_b}+\frac{1}{2}\biggr), \nonumber \\
K^b_{ij} & = &
\frac{\delta_{ij}}{16\pi^2}
\biggl(\frac{1}{4\epsilon_b}\biggr), \nonumber \\
L^b_0 & \equiv & \Tint{p_b}'\frac{p_0^4}{(p^2)^4}=
\frac{1}{16\pi^2}
\biggl(\frac{1}{8\epsilon_b}+\frac{1}{3}\biggr).
\nonumber
\end{eqnarray}
Here we did not write
$\mu^{-2\epsilon}$
explicitly and neglected the higher-order
contributions in~$k^2/T^2$.
For the fermionic case one simply replaces $1/\epsilon_b$
by $1/\epsilon_f$ everywhere in eq.~\nr{integrals}.
\subsubsection{Wave function normalization}
\label{3dwf}
Let us calculate how the 3d fields are
related to the 4d fields. This is to be done on 1-loop level.
In practice, one has to calculate the contribution of the superheavy modes
to the momentum-dependent part of the two-point correlator
of the light and heavy modes. Indeed, the contribution of
the light and heavy modes
is the same in the full theory and the effective theory, whereas
the contribution of the superheavy modes
can be produced in the effective theory
only by a different normalization of the fields.
The generic diagrams needed for the scalar correlator,
and for the temporal and spatial components of the
vector correlator, are shown in Figs.~\ref{drpi}.a and~\ref{drpi}.b.
To determine the wave function normalization factor,
one needs only the parts proportional to $k^2$ from these diagrams.
We identify the diagram by the types of propagators that appear
in it, S, V, F, and $\eta$ denoting the scalar, vector, fermion
and ghost propagators. Counterterm contributions are denoted by CT.
After some simple algebra one gets
for the diagrams of Fig.~\ref{drpi}.a the results
\begin{eqnarray}
{\cal Z}^{\phi}_{\rm CT} & = & k^2 \delta Z_S,\label{zfct}\\
{\cal Z}^{\phi}_{\rm SV} & = &
\Tint{p_b}' \frac{(2 k_\mu+p_\mu)(2k_\nu+p_\nu)
\Bigl(\delta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\Bigr)}
{p^2[(p+k)^2+m_S^2]} \nonumber \\
& \Rightarrow & 4k^2B_b'-4k_ik_j K^b_{ij}
=\frac{k^2}{16\pi^2}\frac{3}{\epsilon_b}, \label{zfsv}\\
{\cal Z}^{\phi}_{\rm FF} & = &
\Tint{p_f} \frac{\mathop{\rm Tr}(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p})[i\@ifnextchar[{\@slash}{\@slash[\z@]}[-0.4mm]{(p+k)}]}
{p^2(p+k)^2} \nonumber \\
& \Rightarrow & 2k^2B_f'=\frac{k^2}{16\pi^2}\frac{2}{\epsilon_f}.
\label{zfff}
\end{eqnarray}
When the correct coefficients are taken into account,
the counterterm contribution
${\cal Z}^{\phi}_{\rm CT}$ cancels the $1/\epsilon$-parts from
the two other contributions,
since there is no wave-function renormalization
in the 3d theory.
The remaining $L_b$- and $L_f$-terms determine the relation
of the 3d fields to the 4d fields.
Explicit expressions for the EW theory are given in Sec.~\ref{DRinSM}.
For the vector correlator, the spatial and temporal components
have to be calculated separately.
For the spatial components, one only needs to calculate the transversal
part, and hence the longitudinal part is not displayed below.
The symbols $J^{b(T)}_{ij}$, $K^{b(T)}_{ij}$ mean
the transversal parts of $J^{b}_{ij}$, $K^{b}_{ij}$ in eq.~\nr{integrals}.
The diagrams in Fig.~\ref{drpi}.b give
\begin{eqnarray}
{\cal Z}^{A_0}_{\rm CT} & = & k^2\delta Z_V,\label{za0ct}\\
& & \nonumber \\
{\cal Z}^{A_i}_{\rm CT} & = & k^2
\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)\delta Z_V,\\
& & \nonumber \\
{\cal Z}^{A_0}_{\rm SS} & = &
\Tint{p_b}'\frac{(2p_0)(2p_0)}{[p^2+m_S^2]
[(p+k)^2+m_S^2]} \nonumber \\
& \Rightarrow & 4J^b_{00}=\frac{k^2}{16\pi^2}
\biggl(-\frac{1}{3\epsilon_b}-\frac{2}{3}\biggr),\label{za0ss}\\
& & \nonumber \\
{\cal Z}^{A_i}_{\rm SS} & = &
\Tint{p_b}'\frac{(2p_i+k_i)(2p_j+k_j)}{[p^2+m_S^2]
[(p+k)^2+m_S^2]} \nonumber \\
& \Rightarrow & 4J^{b(T)}_{ij}=
\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{k^2}{16\pi^2}
\biggl(-\frac{1}{3\epsilon_b}\biggr),\label{zaiss}\\
& & \nonumber \\
{\cal Z}^{A_0}_{\rm \eta\eta} & = &
\Tint{p_b}'\frac{p_0^2}{p^2
(p+k)^2} \nonumber \\
& \Rightarrow & J^{b}_{00}=
\frac{k^2}{16\pi^2}
\biggl(-\frac{1}{12\epsilon_b}-\frac{1}{6}\biggr),\label{za0ee}\\
& & \nonumber \\
{\cal Z}^{A_i}_{\rm \eta\eta} & = &
\Tint{p_b}'\frac{p_i(p_j+k_j)}{p^2
(p+k)^2} \nonumber \\
& \Rightarrow & J^{b(T)}_{ij}=
\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{k^2}{16\pi^2}
\biggl(-\frac{1}{12\epsilon_b}\biggr), \\
& & \nonumber \\
{\cal Z}^{A_0}_{\rm FF} & = &
\Tint{p_f}\frac{\mathop{\rm Tr}
[i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_0 a_L][i\@ifnextchar[{\@slash}{\@slash[\z@]}[-0.4mm]{(p+k)}\gamma_0 a_L]}
{p^2(p+k)^2} \nonumber \\
& = & -2\; \Tint{p_f}\frac{
2p_0^2-\delta_{00}(p^2+p\cdot k)}
{p^2(p+k)^2} \nonumber \\
& \Rightarrow & -4J^f_{00}-k^2B_f'=\frac{k^2}{16\pi^2}
\biggl(-\frac{2}{3\epsilon_f}+\frac{2}{3}\biggr),
\label{za0ff}\\
& & \nonumber \\
{\cal Z}^{A_i}_{\rm FF} & = &
\Tint{p_f}\frac{\mathop{\rm Tr}
[i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_i a_L][i\@ifnextchar[{\@slash}{\@slash[\z@]}[-0.4mm]{(p+k)}\gamma_j a_L]}
{p^2(p+k)^2} \nonumber \\
& = & -2\; \Tint{p_f}\frac{
2p_ip_j+p_ik_j+p_jk_i-\delta_{ij}(p^2+p\cdot k)}
{p^2(p+k)^2} \nonumber \\
& \Rightarrow & -4J^{f(T)}_{ij}-
\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
k^2B_f' =
\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{k^2}{16\pi^2}
\biggl(-\frac{2}{3\epsilon_f}\biggr). \label{zaiee}
\end{eqnarray}
To give the two remaining
contributions ${\cal Z}^{A_0}_{\rm VV}$ and ${\cal Z}^{A_i}_{\rm VV}$,
we note that
\begin{equation}
{\cal Z}^{A_0}_{\rm VV} = \Tint{p_b}'\frac{M_{00}}{p^2(p+k)^2},\quad
{\cal Z}^{A_i}_{\rm VV} = \Tint{p_b}'\frac{M_{ij}}{p^2(p+k)^2},
\end{equation}
where, apart from terms proportional to $k_\mu$,~\cite{PT}
\begin{eqnarray}
M_{\mu\nu}
& = &
\biggl(\delta_{\alpha\beta}-\frac{p_\alpha p_\beta}{p^2}\biggr)
\biggl(\delta_{\sigma\rho}-\frac{(p+k)_\sigma (p+k)_\rho}{(p+k)^2}\biggr)
F_{\mu\alpha\sigma}(k,p,-p-k)F_{\nu\beta\rho}(k,p,-p-k) \nonumber \\
& \Rightarrow & \Bigl[p^2+(p+k)^2+4k^2\Bigr]\delta_{\mu\nu}+
(10-8\epsilon)p_\mu p_\nu \nonumber \\
& & -\frac{2}{p^2}
\Bigl[(p^2+2p\cdot k)^2\delta_{\mu\nu}-
(p^2+2p\cdot k-k^2)p_\mu p_\nu \Bigr] +
\frac{1}{p^2(p+k)^2}\Bigl[k^4p_\mu p_\nu\Bigr].\label{Muv}
\end{eqnarray}
Here we utilized the symmetry of the integrand in the change $p\to -p-k$.
The results for the $k^2$-terms can then be seen to be
\begin{eqnarray}
{\cal Z}^{A_0}_{\rm VV}
& \Rightarrow &
4k^2 B_b'+(10-8\epsilon)J^b_{00}-
2\Bigl[-k^2B_b'+2k^2K^b_{00}\Bigr]=
\frac{k^2}{16\pi^2}
\biggl(\frac{25}{6}\frac{1}{\epsilon_b}-3\biggr),\label{za0vv}\\
& & \nonumber \\
{\cal Z}^{A_i}_{\rm VV}
& \Rightarrow &
4k^2 B_b'\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
+(10-8\epsilon)J^{b(T)}_{ij}-
2\Bigl[-k^2B_b'\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)\nonumber \\
& & +2k^2K^{b(T)}_{ij}\Bigr] =
\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{k^2}{16\pi^2}
\biggl(\frac{25}{6}\frac{1}{\epsilon_b}+\frac{2}{3}\biggr).\label{zaivv}
\end{eqnarray}
The constant $2/3$ in eq.~\nr{zaivv}
comes from $-8\epsilon J^{b(T)}_{ij}$.
When all the contributions are summed together
with the correct coefficients,
the counterterm contributions ${\cal Z}^{A}_{\rm CT}$
again cancel the $1/\epsilon$-parts.
\subsubsection{The couplings of the gauge sector}
\label{3dgs}
To calculate the couplings of the gauge sector,
one has to study some vertex to which
the gauge fields couple. The spatial gauge fields feel only
one coupling constant~$g_3^2$ due to gauge invariance. The interaction
of the temporal components of the gauge fields with the other scalar
fields is not protected by gauge invariance, and hence the
corresponding couplings may differ from $g_3^2$. We calculate
the couplings related to the gauge fields from
a four-point correlator, since
the external momenta may then be assumed to be zero.
In practice, it is most convenient to choose
the $(\phi\phi A A)$-correlator,
since then one gets the two couplings related to
the $(\phi\phi A_iA_j)$- and $(\phi\phi A_0A_0)$-vertices from
almost the same calculations.
The diagrams needed are shown in Fig.~\ref{drg3}.
The results are (${\cal G}_{0}$ is the tree-level contribution)
\begin{eqnarray}
{\cal G}^{A_0}_{0} & = &
{\cal G}^{A_i}_{0} = g^2, \label{ga00}
\\ \nonumber \\
{\cal G}^{A_0}_{\rm CT} & = &
{\cal G}^{A_i}_{\rm CT} = \delta g^2,
\\ \nonumber \\
{\cal G}^{A_0}_{\rm SS} & = &
\Tint{p_b}'\frac{1}{(p^2)^2}=B_b'=\frac{1}{16\pi^2}\frac{1}{\epsilon_b},
\\ \nonumber \\
{\cal G}^{A_i}_{\rm SS} & = &
\Tint{p_b}'\frac{\delta_{ij}}{(p^2)^2}=\delta_{ij}B_b'=
\frac{\delta_{ij}}{16\pi^2}\frac{1}{\epsilon_b}, \label{gaiss}
\\ \nonumber \\
{\cal G}^{A_0}_{\rm SV} & = &
\Tint{p_b}'\frac{
\Bigl(\delta_{00}-\frac{p_0^2}{p^2}\Bigr)}{(p^2)^2}
\nonumber \\
& & = B_b'- K^b_{00}
=\frac{1}{16\pi^2}
\biggl(\frac{3}{4}\frac{1}{\epsilon_b}-\frac{1}{2}\biggr),
\\ \nonumber \\
{\cal G}^{A_i}_{\rm SV} & = &
\Tint{p_b}'\frac{
\Bigl(\delta_{ij}-\frac{p_ip_j}{p^2}\Bigr)}{(p^2)^2}
\nonumber \\
& & = \delta_{ij}B_b'- K^b_{ij}
=\frac{\delta_{ij}}{16\pi^2}
\biggl(\frac{3}{4}\frac{1}{\epsilon_b}\biggr),
\\ \nonumber \\
{\cal G}^{A_0}_{\rm VV} & = &
\Tint{p_b}'\frac{
\Bigl(\delta_{\alpha\mu}-\frac{p_\alpha p_\mu}{p^2}\Bigr)
\Bigl(\delta_{\alpha\nu}-\frac{p_\alpha p_\nu}{p^2}\Bigr)}{(p^2)^2}
\Bigl(2\delta_{\mu\nu}\delta_{00}-2\delta_{\mu 0}\delta_{\nu 0}\Bigr)
\nonumber \\
& & = 4(1-\epsilon) B_b'+2 K^b_{00}
=\frac{1}{16\pi^2}
\biggl(\frac{9}{2}\frac{1}{\epsilon_b}-3\biggr),
\\ \nonumber \\
{\cal G}^{A_i}_{\rm VV} & = &
\Tint{p_b}'\frac{
\Bigl(\delta_{\alpha\mu}-\frac{p_\alpha p_\mu}{p^2}\Bigr)
\Bigl(\delta_{\alpha\nu}-\frac{p_\alpha p_\nu}{p^2}\Bigr)}{(p^2)^2}
\Bigl(2\delta_{\mu\nu}\delta_{ij}-\delta_{\mu i}\delta_{\nu j}
-\delta_{\mu j}\delta_{\nu i}\Bigr)
\nonumber \\
& & = 4(1-\epsilon) B_b'\delta_{ij}+2 K^b_{ij}
=\frac{\delta_{ij}}{16\pi^2}
\biggl(\frac{9}{2}\frac{1}{\epsilon_b}-4\biggr),
\\ \nonumber \\
{\cal G}^{A_0}_{\rm SSS} & = &
\Tint{p_b}'\frac{(2p_0)^2}{(p^2)^3} =
4 K^b_{00}=\frac{1}{16\pi^2}
\biggl(\frac{1}{\epsilon_b}+2\biggr),
\\ \nonumber \\
{\cal G}^{A_i}_{\rm SSS} & = &
\Tint{p_b}'\frac{(2p_i)(2p_j)}{(p^2)^3} =
4 K^b_{ij}=\frac{\delta_{ij}}{16\pi^2}
\biggl(\frac{1}{\epsilon_b}\biggr), \label{gaisss}
\\ \nonumber \\
{\cal G}^{A_0}_{\rm VVV} & = &
\Tint{p_b}'
\frac{
\Bigl(\delta_{\alpha\mu}-\frac{p_\alpha p_\mu}{p^2}\Bigr)
\Bigl(\delta_{\alpha\nu}-\frac{p_\alpha p_\nu}{p^2}\Bigr)
\Bigl(\delta_{\sigma\rho}-\frac{p_\sigma p_\rho}{p^2}\Bigr)}
{(p^2)^3}
F_{0\mu\sigma}(0,p,-p)F_{0\nu\rho}(0,p,-p) \nonumber \\
& = &
4 (3-2\epsilon)\Tint{p_b}'
\frac{p_0^2}{(p^2)^3}=
4 (3-2\epsilon) K^b_{00}=
\frac{1}{16\pi^2}
\biggl(\frac{3}{\epsilon_b}+4\biggr),
\\ \nonumber \\
{\cal G}^{A_i}_{\rm VVV} & = &
\Tint{p_b}'
\frac{
\Bigl(\delta_{\alpha\mu}-\frac{p_\alpha p_\mu}{p^2}\Bigr)
\Bigl(\delta_{\alpha\nu}-\frac{p_\alpha p_\nu}{p^2}\Bigr)
\Bigl(\delta_{\sigma\rho}-\frac{p_\sigma p_\rho}{p^2}\Bigr)}
{(p^2)^3}
F_{i\mu\sigma}(0,p,-p)F_{j\nu\rho}(0,p,-p) \nonumber \\
& = &
4 (3-2\epsilon)\Tint{p_b}'
\frac{p_ip_j}{(p^2)^3}=
4 (3-2\epsilon) K^b_{ij}=
\frac{\delta_{ij}}{16\pi^2}
\biggl(\frac{3}{\epsilon_b}-2\biggr),
\\ \nonumber \\
{\cal G}^{A_0}_{\rm FFFF} & = &
\Tint{p_f}\frac{1}{(p^2)^4}
\mathop{\rm Tr} [(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p})(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p})(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_0 a_L)
(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p} \gamma_0 a_L)] \nonumber \\
& = &
2\;\Tint{p_f}\frac{2p_0^2-p^2\delta_{00}}{(p^2)^3}
=4K^f_{00}-2B_f'
=\frac{1}{16\pi^2}
\biggl(-\frac{1}{\epsilon_f}+2\biggr),
\\ \nonumber \\
{\cal G}^{A_i}_{\rm FFFF} & = &
\Tint{p_f}\frac{1}{(p^2)^4}
\mathop{\rm Tr} [(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p})(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p})(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_i a_L)
(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p} \gamma_j a_L)] \nonumber \\
& = &
2\;\Tint{p_f}\frac{2p_ip_j-p^2\delta_{ij}}{(p^2)^3}
=4K^f_{ij}-2B_f'\delta_{ij}
=\frac{\delta_{ij}}{16\pi^2}
\biggl(-\frac{1}{\epsilon_f}\biggr). \label{gaiffff}
\end{eqnarray}
The counterterm contributions
${\cal G}^{A_0}_{\rm CT}$, ${\cal G}^{A_i}_{\rm CT}$
cancel the $1/\epsilon$-parts from the other
contributions, since there is no
coupling constant renormalization in the 3d theory.
The final result for the 3d couplings
then consists of the tree-level result
corrected by logarithmic terms and constants.
\subsubsection{The couplings of the fundamental scalar sector}
\label{fundam}
To derive the 3d mass and self-coupling of the $\phi$-field,
one has to calculate the effect of the superheavy modes on
the two- and four-point scalar correlators
with vanishing external momenta.
These contributions can most easily be derived by calculating the
effective potential $V(\varphi)$ for the scalar field,
and extracting from it the part coming from the superheavy modes.
The effective potential contains the one-particle-irreducible
Green's functions $G_n$ at vanishing external momenta
through $V(\varphi)=\sum_n(1/n!)G_n\varphi^n$,
so that the terms
quadratic and quartic in $\varphi$ give the two- and four-point correlators.
The usefulness of $V(\varphi)$ lies in the fact
that the combinatorial factors associated with it
are simpler than those associated
with a direct evaluation of Feynman diagrams.
Since
the superheavy modes do not suffer from IR-problems,
their contribution to the effective potential
is analytic in the mass parameters appearing in the
propagators. In contrast to the direct evaluation of superheavy
contributions in the previous sections, the masses cannot here
be neglected, though, but are quite essential:
the mass parameters depend quadratically
on the shifted field $\varphi$, so that terms of the
form $T^2m^2$, $m^4$ determine the
two- and four-point scalar correlators. To get
the quartic coupling, it is enough to extract the $m^4$-term from
the 1-loop effective potential. For the mass
parameter $m_3^2$, however, one needs a 2-loop calculation.
Let us note that to get the correct result to
order $g^4$ for $V(\varphi)$
actually requires resummation~\cite{AE}. This can be done by
adding and subtracting from the Lagrangian the 1-loop
thermal mass terms
\begin{equation}
\overline{\Pi}_{\phi}(0)\phi(0,\vec{k})\phi(0,-\vec{k}),\quad
\frac{1}{2}\overline{\Pi}_{A_0}(0)A_0(0,\vec{k})A_0(0,-\vec{k}),
\label{TCT}
\end{equation}
where the bar indicates that only contributions from
the superheavy modes are included.
The terms added to $\cal L$ with plus-signs are treated as tree-level
masses, whereas the terms subtracted with minus-signs
are treated as counterterms. For the present problem, however,
resummation is inessential, since it affects only
the contributions coming from the $n=0$-modes.
In other words, it is sufficient to know which
contributions to $V(\varphi)$ come from 3d, but
the exact expressions are not needed.
Hence the thermal corrections to masses may be neglected,
allowing one to treat the temporal and spatial components
of the gauge fields as having the same mass,
which simplifies the expressions somewhat.
Just for cosmetic reasons, one might wish to
calculate the 1-loop contributions from the
thermal counterterms, though, since they cancel the
linear terms of the form $mT^3$ in the unresummed
2-loop $V(\varphi)$, see below.
To calculate $V(\varphi)$, one shifts $\phi\to\phi+\varphi$,
neglects linear terms, and calculates all the
one-particle-irreducible vacuum diagrams.
The masses of the
scalar, vector and fermion fields, respectively,
are of the form
\begin{equation}
m^2=m_S^2+n_1\lambda\varphi^2,\quad
M^2=n_2 g^2\varphi^2,\quad
m_f^2=n_3g_Y^2\varphi^2, \label{masses}
\end{equation}
where $n_1$, $n_2$, $n_3$ are some numerical factors.
The propagators in eq.~\nr{symprop} change accordingly.
The ghosts remain massless in Landau gauge.
The shift generates mass counterterms ($\delta m^2$,
$\delta M^2$ and $\delta m_f$) from the corresponding
coupling constant counterterms, as well.
To calculate the 1-loop contribution to $V(\varphi)$,
one needs the integrals~\cite{AE}
\begin{eqnarray}
J_b(m) & = & \frac{1}{2}\Tint{p_b} \ln ({p^2+m^2})=
\mu^{-2\epsilon}
\biggl[\frac{m^2T^2}{24}-\frac{m^3T}{12\pi}-
\frac{m^4}{64\pi^2}\biggl(\frac{1}{\epsilon}+L_b\biggr)
\biggr]+{\cal O}\Bigl(\frac{m^6}{T^2}\Bigr), \nonumber \\
J_f(m) & = & \frac{1}{2}\Tint{p_f} \ln ({p^2+m^2})=\mu^{-2\epsilon}
\biggl[-\frac{m^2T^2}{48}-
\frac{m^4}{64\pi^2}\biggl(\frac{1}{\epsilon}+L_f\biggr)
\biggr]+{\cal O}\Bigl(\frac{m^6}{T^2}\Bigr). \label{JbJf}
\end{eqnarray}
The terms suppressed by $T^2$
and neglected in eq.~\nr{JbJf} are
\begin{equation}
J_b^{(6)}=\frac{\zeta(3)}{768\pi^4}\frac{m^6}{T^2},\quad
J_f^{(6)}=\frac{7\zeta(3)}{768 \pi^4}\frac{m^6}{T^2}
\label{hoJ},
\end{equation}
and give
the higher order operators discussed in Sec.~\ref{corrections}.
In the 3d theory, the integral corresponding
to $J_b(m)$ is
\begin{equation}
J_3(m)=\frac{T}{2}\int \frac{d^dp}{(2\pi)^d}
\ln ({p^2+m^2})=\mu^{-2\epsilon}
\bigg(-\frac{m^3T}{12\pi}\biggr),
\end{equation}
which is just a part of $J_b(m)$. With these integrals, one can write
the typical scalar, vector and fermion contributions
${\cal C}_S(m)$, ${\cal C}_V(M)$ and ${\cal C}_F(m_f)$
to $V_1(\varphi)$, and separate from these the 3d-part.
The massless ghosts do not contribute. The results are
\begin{eqnarray}
{\cal C}_S(m) & \equiv &
-\Tint{p_b}\ln \biggl(\frac{1}{p^2+m^2}\biggr)^{1/2}=J_b(m) \nonumber \\
& = & {\cal C}_S^{3d}+
\mu^{-2\epsilon}
\biggl[\frac{m^2T^2}{24}-
\frac{m^4}{64\pi^2}\biggl(\frac{1}{\epsilon}+L_b\biggr)
\biggr],\label{cs} \\
& & \nonumber \\
{\cal C}_V(M) & \equiv &
-\Tint{p_b}\ln \biggl(
\det \frac{\delta_{\mu\nu}-p_\mu p_\nu/p^2}{p^2+M^2}
\biggr)^{1/2}=(3-2\epsilon)J_b(M) \nonumber \\
& = & {\cal C}_V^{3d}+
\mu^{-2\epsilon}
\biggl[\frac{M^2T^2}{8}-
\frac{M^4}{64\pi^2}\biggl(\frac{3}{\epsilon}+3L_b-2\biggr)
\biggr], \\
& & \nonumber \\
{\cal C}_F(m_f) & \equiv &
\Tint{p_f}\ln \det\frac{1}{i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}+m_f}=-4 J_f(m_f) \nonumber \\
& = &
\mu^{-2\epsilon}
\biggl[\frac{m_f^2T^2}{12}+
\frac{m_f^4}{16\pi^2}\biggl(\frac{1}{\epsilon}+L_f\biggr)
\biggr],\label{cf}
\end{eqnarray}
where ${\cal C}_S^{3d}$ and ${\cal C}_V^{3d}$
are the corresponding integrals in the 3d theory.
The $1/\epsilon$-parts are $T$-independent, and are
cancelled by the 1-loop counterterms
$\frac{1}{2}\delta m_S^2\varphi^2$, $\frac{1}{4}\delta \lambda\varphi^4$.
Since the bosonic field content of the 3d theory is the same
as that of the original theory, the parts
${\cal C}_S^{3d}$ and ${\cal C}_V^{3d}$
are reproduced by the 3d theory.
The coefficient of $\varphi^2/2$ of the
remaining terms determines the 1-loop
result for $m_3^2$, and the coefficient of $\varphi^4/4$
the 1-loop result for $\lambda_3$.
In this simple way,
the couplings of the fundamental scalar sector
in the effective 3d theory get fixed at 1-loop order.
Next we go to 2-loop level, which is
required for the mass $m_3^2$.
In general, there are three classes of
diagrams (see, e.g., Fig.~23 in~\cite{AE}) contributing
at order $g^4$: the sunset diagrams,
the figure~8 -diagrams, and
the 1-loop counterterm diagrams.
The counterterm diagrams can contain either
the mass or the wave function counterterm.
The general strategy is the same as at 1-loop level: from
each bosonic diagram, one separates the contribution
coming from the $n=0$ modes, since this
contribution is reproduced by the 3d theory.
The remaining part, analytic in the mass parameters,
is not reproduced by the 2-loop diagrams of the 3d theory,
and must hence be due to corrections to the tree-level
parameters of the 3d theory. The fermionic diagrams
do not appear in the 3d theory,
but they are IR-safe, and hence directly produce
terms analytic in the mass parameters, contributing to $m_3^2$.
We will first give the basic integrals appearing
in the calculation, and then the results
for the contributions of the superheavy modes to
all the different types of diagrams that can appear.
The bosonic tadpole integral is
\begin{equation}
I_b(m)=\Tint{p_b} \frac{1}{p^2+m^2}=I_b'(m)+I_3(m),
\end{equation}
where $I'_b(m)$ is in eq.~\nr{Ib}
and the 3d integral is
\begin{equation}
I_3(m)=T\int \frac{d^dp}{(2\pi)^d}\frac{1}{p^2+m^2}=\mu^{-2\epsilon}
\bigg(-\frac{mT}{4\pi}\biggr). \label{I3}
\end{equation}
The fermionic tadpole integral
$I_f(m)$ is given in eq.~\nr{If}.
The products appearing in the 2-loop diagrams,
apart from inessential vacuum terms, are
\begin{eqnarray}
I_b(m_1)I_b(m_2) & = & I_3(m_1)I_3(m_2)
+\frac{1}{12}\mu^{-2\epsilon}T^2\Bigl[I_3(m_{1})+I_3(m_{2})\Bigr]\nonumber \\
& & -\mu^{-4\epsilon}\frac{T^2}{16\pi^2}
(m_{1}^2+m_{2}^2)\biggl(
\frac{1}{12\epsilon}+
\frac{L_b}{12}+\frac{\imath_\epsilon}{12}\biggr), \label{f2l} \\
I_b(m)I_f(m_f) & = &
-\frac{1}{24}\mu^{-2\epsilon}T^2 I_3(m)
+ \mu^{-4\epsilon}\frac{T^2}{16\pi^2}\biggl[
m^2\biggl(
\frac{1}{24\epsilon}+
\frac{L_b}{24}+\frac{\imath_\epsilon}{24}-\frac{1}{12}\ln 2\biggr),
\nonumber \\
& &
-m_f^2\biggl(
\frac{1}{12\epsilon}+
\frac{L_f}{12}+\frac{\imath_\epsilon}{12}\biggr)\biggr], \\
I_f(m_f)I_f(m_{f'}) & = &
\mu^{-4\epsilon}\frac{T^2}{16\pi^2}
(m_f^2+m_{f'}^2)\biggl(
\frac{1}{24\epsilon}+
\frac{L_f}{24}+\frac{\imath_\epsilon}{24}-\frac{1}{12}\ln 2\biggr).
\end{eqnarray}
The bosonic sunset integral is~\cite{FKRS1,AZ}
\begin{eqnarray}
H_b(m_1,m_2,m_3) & = &\Tint{p_b,q_b}\frac{1}
{[p^2+m_1^2][q^2+m_2^2][(p+q)^2+m_3^2]} \nonumber \\
& = & \mu^{-4\epsilon}
\frac{T^2}{16\pi^2}\biggl(
\frac{1}{4\epsilon}+\ln\frac{\mu}{m_1+m_2+m_3}+\frac{1}{2}\biggr) \nonumber \\
& = & H_3(m_1,m_2,m_3)+\mu^{-4\epsilon}
\frac{T^2}{16\pi^2}\biggl(
\frac{1}{4}L_b+\frac{1}{4}\imath_\epsilon+
\ln\frac{3T}{\mu}+c\biggr),\label{l2l}
\end{eqnarray}
where $H_3(m_1,m_2,m_3)$ is the corresponding integral
in 3d. Using eq.~\nr{nabi} one can see
that the expression in the brackets
on the last line actually vanishes,
so that $H_3(m_1,m_2,m_3)$ is the second
line in eq.~\nr{l2l}.
However, it proves useful to write
$H_b(m_1,m_2,m_3)$ in the form indicated.
The fermionic integral
\begin{eqnarray}
H_f(m_f,m_{f'},m) & = &\Tint{p_f,q_f}\frac{1}
{[p^2+m_f^2][q^2+m_{f'}^2][(p+q)^2+m^2]}
\end{eqnarray}
vanishes~\cite{AE}.
There is also a third integral, $L(m_1,m_2)$, appearing in the
2-loop graphs~\cite{AE}.
Apart from vacuum terms, it is given by
\begin{eqnarray}
L(m_1,m_2) & = &
\Tint{p_b,q_b}\frac{(p\cdot q)^2}{p^2(p^2+m_{1}^2)
q^2(q^2+m_{2}^2)}
\nonumber \\
& = & L_3(m_1,m_2)
+\mu^{-2\epsilon}\frac{T^2}{24}\Bigl[I_3(m_{1})+I_3(m_{2})\Bigr]
\nonumber \\
& - & \mu^{-4\epsilon}\frac{T^2}{16\pi^2}
(m_{1}^2+m_{2}^2)
\biggl(
\frac{1}{48\epsilon}+
\frac{L_b}{48}+\frac{\imath_\epsilon}{48}-\frac{1}{48}\biggr),
\end{eqnarray}
where
\begin{equation}
L_3(m_1,m_2) = \frac{1}{3}I_3(m_1)I_3(m_2).
\end{equation}
However, this integral does not contribute
to the integration over the superheavy scale
in the Standard Model,
since it is cancelled between the figure~8 and sunset
diagrams containing only SU(2)
vector fields (${\cal D}_{\rm VV}$
and ${\cal D}_{\rm VVV}$ below)~\cite{AE}.
Using the given integrals and results from~\cite{AE}
for the 2-loop diagrams, one can write down
the contributions from the superheavy modes to all the
possible types of 2-loop diagrams.
We give here explicit results only for
the simplest mass combinations in the propagators,
relevant for Sec.~\ref{DRinSM};
the results for the cases with other masses
can be read by using eqs.~\nr{f2l}-\nr{l2l} and Appendix A of~\cite{AE}.
As stated above, we need not bother about resummation.
The results are
\begin{eqnarray}
{\cal D}_{\rm SSS}(m_{1},m_{2},m_{3}) & = &
\Tint{p_b,q_b}\frac{1}
{[p^2+m_{1}^2][q^2+m_{2}^2][(p+q)^2+m_{3}^2]} \nonumber \\
& = & {\cal D}^{3d}_{\rm SSS}+
\mu^{-4\epsilon}\frac{T^2}{16\pi^2}\biggl(
\frac{L_b}{4}+\frac{\imath_\epsilon}{4}+\ln\frac{3T}{\mu}+c\biggr),
\label{fae} \\
& & \nonumber \\
{\cal D}_{\rm SSV}(m_{1},m_{2},M) & = &
\Tint{p_b,q_b}\frac{(2p_\mu+q_\mu)(2p_\nu+q_\nu)\Bigl(\delta_{\mu\nu}-
\frac{q_\mu q_\nu}{q^2}\Bigr)}
{[p^2+m_{1}^2][q^2+M^2][(p+q)^2+m_{2}^2]}
\nonumber \\
& = & {\cal D}^{3d}_{\rm SSV}+
\frac{1}{6}\mu^{-2\epsilon}T^2I_3(M) \nonumber \\
& &
+\mu^{-4\epsilon}\frac{T^2}{16\pi^2}\biggl[
M^2 \biggl(
-\frac{1}{6\epsilon}+
\frac{L_b}{12}+\frac{\imath_\epsilon}{12}+\ln\frac{3T}{\mu}+c
\biggr) \nonumber \\
& &
-2(m_{1}^2+m_{2}^2) \biggl(
\frac{L_b}{4}+\frac{\imath_\epsilon}{4}+\ln\frac{3T}{\mu}+c
\biggr)\biggr], \\
& & \nonumber \\
{\cal D}_{\rm SVV}(m,M,M) & = &
\Tint{p_b,q_b}\frac{4\Bigl(\delta_{\mu\nu}-
\frac{p_\mu p_\nu}{p^2}\Bigr)
\Bigl(\delta_{\mu\nu}-\frac{q_\mu q_\nu}{q^2}\Bigr)}
{[p^2+M^2][q^2+M^2][(p+q)^2+m^2]}
\nonumber \\
& = & {\cal D}^{3d}_{\rm SVV}+
\mu^{-4\epsilon}\frac{T^2}{16\pi^2}\biggl[
\frac{5}{2}L_b+\frac{5}{2}\imath_\epsilon+
10\biggl(\ln\frac{3T}{\mu}+c\biggr)
\biggr], \\
& & \nonumber \\
{\cal D}_{\rm VVV}(M,M,M) & = & \!\!\!\!
\Tint{p_b,q_b}\!\!\!\frac{
\Bigl(\delta_{\mu\alpha}-\frac{p_\mu p_{\alpha}}{p^2}\Bigr)
\Bigl(\delta_{\nu\beta}-\frac{q_\nu q_{\beta}}{q^2}\Bigr)
\Bigl(\delta_{\rho\gamma}-\frac{r_\rho r_{\gamma}}{r^2}\Bigr)}
{[p^2+M^2][q^2+M^2][(p+q)^2+M^2]}
F_{\mu\nu\rho}(p,q,r)F_{\alpha\beta\gamma}(p,q,r)
\nonumber \\
& = & \tilde{\cal D}^{3d}_{\rm VVV}+\frac{7}{4}\mu^{-2\epsilon}T^2I_3(M)
-3L(M,M) \nonumber \\
& &
-\mu^{-4\epsilon}\frac{T^2}{16\pi^2}M^2 \biggl[
\frac{7}{4}\frac{1}{\epsilon}+
\frac{31}{4}L_b+\frac{31}{4}\imath_\epsilon+
24\biggl(\ln\frac{3T}{\mu}+c\biggr)-1\biggr], \\
& & \nonumber \\
{\cal D}_{\rm \eta\eta V}(M) & = &-
\Tint{p_b,q_b}\frac{p_\mu(p_\nu+q_\nu)\Bigl(\delta_{\mu\nu}-
\frac{q_\mu q_\nu}{q^2}\Bigr)}
{p^2(q^2+M^2)(p+q)^2}
\nonumber \\
& = & {\cal D}^{3d}_{\rm \eta\eta V}-
\frac{1}{24}\mu^{-2\epsilon}T^2 I_3(M) \nonumber \\
& &
-\mu^{-4\epsilon}\frac{T^2}{16\pi^2}M^2\biggl[
-\frac{1}{24\epsilon}+
\frac{L_b}{48}+\frac{\imath_\epsilon}{48}+
\frac{1}{4}\biggl(\ln\frac{3T}{\mu}+c\biggr)\biggr], \\
& & \nonumber \\
{\cal D}_{\rm FFS}(m_{f},m_{f'},m) & = &
\Tint{p_f,q_f}\frac{\mathop{\rm Tr}(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}+m_{f})
(i\@ifnextchar[{\@slash}{\@slash[\z@]}{q}-m_{f'})}
{[p^2+m_{f}^2][q^2+m_{f'}^2][(p+q)^2+m^2]}
\nonumber \\
& = & -\frac{1}{6}\mu^{-2\epsilon}T^2I_3(m)+
\mu^{-4\epsilon}\frac{T^2}{16\pi^2}\biggl[
m^2\biggl(\frac{1}{6\epsilon}+
\frac{L_b}{6}+\frac{\imath_\epsilon}{6}-\frac{1}{3}\ln 2\biggr)
\nonumber \\
& &
-(m_{f}^2+m_{f'}^2)\biggl(\frac{1}{4\epsilon}+
\frac{L_f}{4}+\frac{\imath_\epsilon}{4}-\frac{1}{6}\ln 2\biggr)
\biggr], \\
& & \nonumber \\
{\cal D}_{\rm FFV}(m_{f},m_{f'},M) & = &
-\Tint{p_f,q_f}\frac{\mathop{\rm Tr}(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}+m_{f})
\gamma_\mu a_L(i\@ifnextchar[{\@slash}{\@slash[\z@]}{q}-m_{f'})\gamma_\nu a_L}
{[p^2+m_{f}^2][q^2+m_{f'}^2][(p+q)^2+M^2]}
\biggl(\delta_{\mu\nu}-\frac{
r_\mu r_\nu}{r^2}\biggr)
\nonumber \\
& = &
-\frac{1}{6}\mu^{-2\epsilon}T^2I_3(M)+
\mu^{-4\epsilon}\frac{T^2}{16\pi^2}\biggl[
M^2\biggl(\frac{1}{6\epsilon}+
\frac{L_b}{6}+\frac{\imath_\epsilon}{6}-\frac{1}{3}\ln 2-\frac{1}{6}\biggr)
\nonumber \\
& &
-(m_{f}^2+m_{f'}^2)\biggl(\frac{1}{4\epsilon}+
\frac{L_f}{4}+\frac{\imath_\epsilon}{4}-\frac{1}{6}\ln 2-\frac{1}{4}\biggr)
\biggr], \\
& & \nonumber \\
{\cal D}_{\rm SS}(m_{1},m_{2}) & = &
-\Tint{p_b,q_b}\frac{1}{(p^2+m_{1}^2)(q^2+m_{2}^2)}
\nonumber \\
& = & {\cal D}^{3d}_{\rm SS}
-\frac{1}{12}\mu^{-2\epsilon}T^2\Bigl[I_3(m_{1})+I_3(m_{2})\Bigr]\nonumber \\
& & +\mu^{-4\epsilon}\frac{T^2}{16\pi^2}
(m_{1}^2+m_{2}^2)
\biggl(
\frac{1}{12\epsilon}+
\frac{L_b}{12}+\frac{\imath_\epsilon}{12}\biggr), \\
& & \nonumber \\
{\cal D}_{\rm SV}(m,M) & = &
-2\;\Tint{p_b,q_b}\frac{\delta_{\mu\nu}
\Bigl(\delta_{\mu\nu}-\frac{q_\mu q_\nu}{q^2}\Bigr)}
{(p^2+m^2)(q^2+M^2)}
\nonumber \\
& = & {\cal D}^{3d}_{\rm SV}
-\frac{1}{2}\mu^{-2\epsilon}T^2\Bigl[I_3(m)+I_3(M)\Bigr] \nonumber \\
& & +\mu^{-4\epsilon}\frac{T^2}{16\pi^2}
(m^2+M^2)
\biggl(
\frac{1}{2\epsilon}+
\frac{L_b}{2}+\frac{\imath_\epsilon}{2}-\frac{1}{3}\biggr), \\
& & \nonumber \\
{\cal D}_{\rm VV}(M_{1},M_{2}) & = &
-\Tint{p_b,q_b}\frac{\Bigl(\delta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\Bigr)
\Bigl(\delta_{\sigma\rho}-\frac{q_\sigma q_\rho}{q^2}\Bigr)}
{(p^2+M_{1}^2)(q^2+M_{2}^2)}
(2\delta_{\mu\nu}\delta_{\sigma\rho}-
\delta_{\mu\sigma}\delta_{\nu\rho}-
\delta_{\mu\rho}\delta_{\nu\sigma})
\nonumber \\
& = & \tilde{\cal D}^{3d}_{\rm VV}
-\frac{7}{6}\mu^{-2\epsilon}T^2\Bigl[I_3(M_{1})+I_3(M_{2})\Bigr]+
2L(M_{1},M_{2}) \nonumber \\
& & +\mu^{-4\epsilon}\frac{T^2}{16\pi^2}
(M_{1}^2+M_{2}^2)
\biggl(
\frac{7}{6\epsilon}+
\frac{7}{6}L_b+\frac{7}{6}\imath_\epsilon-\frac{5}{3}\biggr), \\
& & \nonumber \\
{\cal D}_S(m) & = &
\Tint{p_b}\frac{\delta m^2+\delta Z_S p^2}{p^2+m^2}
=\mu^{-2\epsilon}\frac{T^2}{12}(1+\epsilon \imath_\epsilon)
(\delta m^2-m^2\delta Z_S),\\
& & \nonumber \\
{\cal D}_V(M) & = &
\Tint{p_b}\frac{\delta M^2+\delta Z_V p^2}{p^2+M^2}
\delta_{\mu\nu}\Bigl(\delta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\Bigr) \nonumber \\
& = & \mu^{-2\epsilon}\frac{T^2}{12}
[3+3\epsilon\imath_\epsilon-2\epsilon]
(\delta M^2-M^2\delta Z_V),\\
& & \nonumber \\
{\cal D}_F(m_f) & = &
\Tint{p_f}\frac{1}{p^2+m_f^2}
\mathop{\rm Tr} [i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}-m_f]
[\delta m_f+i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}(a_L\delta Z^L_F+a_R\delta Z^R_F)] \nonumber \\
& = & \mu^{-2\epsilon}\frac{T^2}{12}
[1+\epsilon (\imath_\epsilon-2\ln 2)]
[\delta m_f^2-m_f^2(\delta Z^L_F+\delta Z^R_F)]. \label{lae}
\end{eqnarray}
In the expressions for
${\cal D}_{\rm VVV}$ and ${\cal D}_{\rm FFV}$,
we used $r$ as a shorthand for $-p-q$.
The tilde over $\tilde{\cal D}^{3d}_{\rm VVV}$ and
$\tilde{\cal D}^{3d}_{\rm VV}$ indicates
that the 3d-part of $L(M,M)$ is not included.
In eq.~\nr{lae}, $\delta m_f^2 \equiv 2 m_f \delta m_f$.
In addition to the diagrams mentioned above,
one may calculate the thermal counterterm diagrams.
According to eq.~\nr{TCT}, they give
contributions of the form
\begin{equation}
-\overline{\Pi}_{\phi}(0)I_3(m),\quad
-\overline{\Pi}_{A_0}(0)I_3(M). \label{thctc}
\end{equation}
These contributions cancel the linear terms proportional
to $T^2I_3(m)$, $T^2I_3(M)$
in eqs.~\nr{fae}-\nr{lae}.
After this cancellation, all that is left
is terms proportional to the masses squared.
Using eq.~\nr{masses},
such terms give the coefficient of $\varphi^2/2$, i.e.,
the mass $m_3^2$ of the 3d theory.
The counterterm contributions
${\cal D}_S(m)$, ${\cal D}_V(M)$, ${\cal D}_F(m_f)$
do not cancel all the $1/\epsilon$-parts from the
other contributions, since there remains a 2-loop mass
counterterm in the 3d theory~\cite{FKRS1}.
\subsubsection{The couplings of the adjoint scalar sector}
\label{adjoint}
The couplings of the adjoint scalar sector
could be calculated from an effective potential
as for the fundamental scalar sector, but
for completeness we calculate them directly from Feynman diagrams.
We make here just a 1-loop calculation.
The 1-loop diagrams needed for calculating
the $(A_0A_0)$-correlator
at vanishing external momenta are shown in Fig.~\ref{drmd}.
We need two new integrals, obtained
by taking the derivative of eqs.~\nr{Ib}, \nr{If}
with respect to~$T$. With the accuracy needed,
\begin{eqnarray}
E_b'(m) & = & \Tint{p_b}'\frac{p_0^2}{(p^2+m^2)^2}=
-\frac{T^2}{24}-
\frac{m^2}{16\pi^2}\biggl(\frac{1}{2\epsilon_b}+1\biggr)
,\nonumber \\
E_f(m) & = &\Tint{p_f}\frac{p_0^2}{(p^2+m^2)^2}=
\frac{T^2}{48}-
\frac{m^2}{16\pi^2}\biggl(\frac{1}{2\epsilon_f}+1\biggr).
\label{Es}
\end{eqnarray}
Note that $E_b'(0)$ is the constant part
subtracted in the definition of $J^b_{00}$
in eq.~\nr{integrals}.
Using the integrals in eq.~\nr{Es} together with those in
eqs.~\nr{Ib},~\nr{integrals}, the diagrams in Fig.~\ref{drmd} give
\begin{eqnarray}
{\cal A}^{(2)}_{\rm S} & = &
\Tint{p_b}'\frac{1}{p^2+m_S^2} = I_b'(m_S) =\frac{T^2}{12}
-\frac{m_S^2}{16\pi^2}\frac{1}{\epsilon_b}, \label{a2s}
\\ & & \nonumber \\
{\cal A}^{(2)}_{\rm V} & = &
\Tint{p_b}'\frac{\delta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}}{p^2}
\Bigl(2\delta_{00}\delta_{\mu\nu}-2\delta_{0\mu}\delta_{0\nu}\Bigr)
= 4(1-\epsilon)I_b'(0)+2E_b'(0) =\frac{T^2}{4},
\\ & & \nonumber \\
{\cal A}^{(2)}_{\rm SS} & = &
\Tint{p_b}'\frac{(2p_0)^2}{(p^2+m_S^2)^2}
= 4E_b'(m_S) =-\frac{T^2}{6}-\frac{m_S^2}{16\pi^2}
\biggl(\frac{2}{\epsilon_b}+4\biggr),
\\ & & \nonumber \\
{\cal A}^{(2)}_{\rm VV} & = &
\Tint{p_b}'\frac{(12-8\epsilon)p_0^2}{(p^2)^2}
= (12-8\epsilon)E_b'(0) =-\frac{T^2}{2},
\\ & & \nonumber \\
{\cal A}^{(2)}_{\eta\eta} & = &
\Tint{p_b}'\frac{p_0^2}{(p^2)^2}
= E_b'(0) =-\frac{T^2}{24},
\\ & & \nonumber \\
{\cal A}^{(2)}_{\rm FF} & = &
-2\Tint{p_f}\frac{2p_0^2-p^2}{(p^2)^2}
= -4E_f(0)+2I_f(0) =-\frac{T^2}{6}. \label{a2ff}
\end{eqnarray}
For ${\cal A}^{(2)}_{\rm SS}$,
${\cal A}^{(2)}_{\rm VV}$,
${\cal A}^{(2)}_{\rm \eta\eta}$, and
${\cal A}^{(2)}_{\rm FF}$,
the integrand was obtained from eqs.~\nr{za0ss}, \nr{Muv},
\nr{za0ee}, and \nr{za0ff} by putting $k\to 0$.
In the final result, the $1/\epsilon$-parts cancel
between ${\cal A}^{(2)}_{\rm S}$ and ${\cal A}^{(2)}_{\rm SS}$.
The constant term proportional to $m_S^2$
is of higher order, and the $T^2$-terms
give the desired 3d mass parameter.
The diagrams needed for the
$(A_0A_0A_0A_0)$-correlator are shown in Fig.~\ref{drla}.
They give
\begin{eqnarray}
{\cal A}^{(4)}_{\rm SS} & = &
\Tint{p_b}'
\frac{1}{(p^2)^2}
=B_b'=
\frac{1}{16\pi^2}
\frac{1}{\epsilon_b}, \label{a4ss}
\\ & & \nonumber \\
{\cal A}^{(4)}_{\rm SSS} & = &
\Tint{p_b}'
\frac{(2p_0)^2}{(p^2)^3}
=4K^b_{00}=
\frac{1}{16\pi^2}
\biggl(\frac{1}{\epsilon_b}+2\biggr),
\\ & & \nonumber \\
{\cal A}^{(4)}_{\rm SSSS} & = &
\Tint{p_b}'
\frac{(2p_0)^4}{(p^2)^4}
=16L^b_0=
\frac{1}{16\pi^2}
\biggl(\frac{2}{\epsilon_b}+\frac{16}{3}\biggr),
\\ & & \nonumber \\
{\cal A}^{(4)}_{\rm VV} & = &
\Tint{p_b}'
\frac{
\Bigl(\delta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\Bigr)
\Bigl(\delta_{\sigma\rho}-\frac{p_\sigma p_\rho}{p^2}\Bigr)}{(p^2)^2}
\nonumber \\
& &
\nonumber \\
& & \times
\Bigl(2\delta_{00}\delta_{\mu\sigma}-
2\delta_{0\mu}\delta_{0\sigma}\Bigr)
\Bigl(2\delta_{00}\delta_{\nu\rho}-
2\delta_{0\nu}\delta_{0\rho}\Bigr)
\nonumber \\
& = &
8 (1-\epsilon) B_b'+
4L^b_0=
\frac{1}{16\pi^2}
\biggl(\frac{17}{2}\frac{1}{\epsilon_b}-\frac{20}{3}\biggr),
\\ & & \nonumber \\
{\cal A}^{(4)}_{\rm VVV} & = &
\Tint{p_b}'
\frac{
\Bigl(\delta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\Bigr)
\Bigl(\delta_{\sigma\alpha}-\frac{p_\sigma p_\alpha}{p^2}\Bigr)
\Bigl(\delta_{\rho\beta}-\frac{p_\rho p_\beta}{p^2}\Bigr)}{(p^2)^3}
\nonumber \\
& &
\nonumber \\
& & \times
F_{0\mu\sigma}(0,p,-p)
F_{0\nu\rho}(0,p,-p)
\Bigl(2\delta_{00}\delta_{\alpha\beta}-
2\delta_{0\alpha}\delta_{0\beta}\Bigr)
\nonumber \\
& = &
16 (1-\epsilon) K^b_{00}+
8L^b_0=
\frac{1}{16\pi^2}
\biggl(\frac{5}{\epsilon_b}+\frac{20}{3}\biggr),
\\ & & \nonumber \\
{\cal A}^{(4)}_{\rm VVVV} & = &
\Tint{p_b}'
\frac{
\Bigl(\delta_{\alpha\beta}-\frac{p_\alpha p_\beta}{p^2}\Bigr)
\Bigl(\delta_{\gamma\delta}-\frac{p_\gamma p_\delta}{p^2}\Bigr)
\Bigl(\delta_{\mu\nu}-\frac{p_\mu p_\nu}{p^2}\Bigr)
\Bigl(\delta_{\sigma\rho}-\frac{p_\sigma p_\rho}{p^2}\Bigr)}{(p^2)^4}
\nonumber \\
& &
\nonumber \\
& & \times
F_{0\alpha\gamma}(0,p,-p)
F_{0\delta\nu}(0,p,-p)
F_{0\mu\sigma}(0,p,-p)
F_{0\beta\rho}(0,p,-p) \nonumber \\
& = &
16 (3-2\epsilon) L^b_0=
\frac{1}{16\pi^2}
\biggl(\frac{6}{\epsilon_b}+12\biggr),
\\ & & \nonumber \\
{\cal A}^{(4)}_{\rm \eta\eta\eta\eta} & = &
\Tint{p_b}'
\frac{p_0^4}{(p^2)^4}
=L^b_0=
\frac{1}{16\pi^2}
\biggl(\frac{1}{8\epsilon_b}+\frac{1}{3}\biggr),
\\ & & \nonumber \\
{\cal A}^{(4)}_{\rm FFFF} & = &
\Tint{p_f}\frac{1}{(p^2)^4}
\mathop{\rm Tr} [(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_0 a_L)(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_0 a_L)
(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_0 a_L)(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_0 a_L)] \nonumber \\
& = &
=16L^f_0-16K^f_{00}+2B_f'
=\frac{1}{16\pi^2}
\biggl(-\frac{8}{3}\biggr). \label{a4ffff}
\end{eqnarray}
Again the $1/\epsilon$-parts cancel
when the correct coefficients are included,
since there are no coupling constant
counterterms in the 3d theory.
\subsection{Integration over the heavy scale}
The integration over the heavy scale proceeds analogously
to the integration over the superheavy scale.
Instead of an infinite number of excitations
with masses $\sim \pi T$, there are
now a finite number of fields with
masses $m_D\sim gT$. The heavy fields include, in particular,
the temporal components $A_0$ of the gauge fields.
The heavy fields are scalars
in the effective 3d theory, so
no gauge fixing is needed for the integration.
The general form of the resulting theory
differs from the starting point
only by the absence of the heavy fields,
so that the final theory is the light bosonic
sector of the original 4d theory,
but in three spatial dimensions.
There are three kinds of vertices which
the heavy field $A_0$ feels.
The interactions with the spatial gauge fields
follow from eq.~\nr{gaugevs}, and are of the form
\begin{equation}
ig_3f^{abc}A_0^a(p)A_i^b(q)A_0^c(r)(p_i-r_i),\quad
g_3^2 G^{abcd}_{ij00}A^a_i A^b_j A^c_0A^d_0,
\label{a0ai}
\end{equation}
where $G^{abcd}_{ij00}\propto\delta_{ij}$.
Here $g_3$ denotes the dimensionful
3d gauge coupling, and the fields have also
been scaled by $T$ to have the dimension GeV$^{1/2}$.
In addition to eq.~\nr{a0ai} there are different
quartic scalar vertices. One of the scalar vertices,
the quartic self-interaction of $A_0$, can be neglected,
since the coupling constant is of order $g^4$,
and would appear only inside 2-loop graphs
where it is further suppressed by other couplings.
The integration measure in 3d is denoted by
\begin{equation}
\int dp \equiv \int\frac{d^dp}{(2\pi)^d}.
\end{equation}
The $A_0$-propagator in the symmetric phase is
\begin{equation}
\langle A_0(p)A_0(-p)\rangle=
\frac{1}{p^2+m_D^2}.
\end{equation}
The basic integrals needed are $\bar{I}_3(m)$, which
is just eq.~\nr{I3} without $T$, together with
${\cal B}(k^2;m_1,m_2)$ and ${\cal J}_{ij}$, defined by
\begin{eqnarray}
{\cal B}(k^2;m_1,m_2) & = &
\int dp\frac{1}{[p^2+m_1^2][(p+k)^2+m_2^2]}\nonumber \\
& = &
\frac{i}{8\pi (k^2)^{1/2}}
\ln\frac{m_1+m_2-i(k^2)^{1/2}}{m_1+m_2+i(k^2)^{1/2}}
\nonumber \\
& = &
\frac{1}{4\pi(m_1+m_2)}\biggl[
1-\frac{1}{3}\frac{k^2}{(m_1+m_2)^2}+\ldots
\biggr],
\\ && \nonumber \\
{\cal J}_{ij} & = &
\int dp\frac{p_ip_j}{[p^2+m_D^2][(p+k)^2+m_D^2]}-
\int dp\frac{p_ip_j}{(p^2+m_D^2)^2}\nonumber \\
& = &
-\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{k^2}{96\pi m_D}+
\frac{k_ik_j}{32 \pi m_D}+{\cal O}\Bigl(\frac{k^4}{m_D^3}\Bigr).
\end{eqnarray}
Using these integrals, the task is to calculate
the corrections from $m_D$ to the parameters of
the final theory in powers of $g_3^2/m_D$, where
it is assumed that the light masses and momenta
are $m\sim k\sim g_3^2$.
Let us start with wave function normalization.
Since the fundamental scalar field interacts
with $A_0$ only through quartic vertices,
there is no momentum-dependent correction from
the $A_0$-field to the $\phi$-correlator.
Hence the normalization of $\phi$ does not change.
The momentum-dependent correction to the $A_i$-correlator
comes from the diagram with two internal $A_0$-propagators,
and is, in analogy with ${\cal Z}^{A_i}_{\rm SS}$
in eq.~\nr{zaiss},
\begin{eqnarray}
{\cal Z}^{A_i}_{\rm LL} & = &
\int dp\frac{(2p_i+k_i)(2p_j+k_j)}
{[p^2+m_D^2][(p+k)^2+m_D^2]}\nonumber \\
& \Rightarrow & 4{\cal J}^{(T)}_{ij}
=-\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{k^2}{24\pi m_D}. \label{zaill}
\end{eqnarray}
The heavy propagators are denoted by L.
To fix the gauge coupling, one can calculate
the contributions of $A_0$ to the $(\phi\phi A_iA_j)$-vertex.
There are two types of diagrams, in analogy with
${\cal G}^{A_i}_{\rm SS}$ and ${\cal G}^{A_i}_{\rm SSS}$
in eqs.~\nr{gaiss}, \nr{gaisss}.
The contributions are
\begin{eqnarray}
{\cal G}_{\rm LL} & = & \int dp \frac{\delta_{ij}}{(p^2+m_D^2)^2}=
\frac{\delta_{ij}}{8\pi m_D}, \label{gll} \\
{\cal G}_{\rm LLL} & = &
\int dp \frac{(2p_i)(2p_j)}{(p^2+m_D^2)^3}=
\frac{\delta_{ij}}{8\pi m_D}. \label{glll}
\end{eqnarray}
It is very easy to calculate the 1-loop
corrections to the parameters of the scalar sector,
since the field one is integrating over is
itself a scalar. Hence the diagrams can give
just $\bar{I}_3(m_D)$ or ${\cal B}(0;m_D,m_D)$.
For the mass parameter, one needs again a 2-loop
calculation, and it is best to employ
the effective potential. After the shift,
the mass of the heavy field is of
the form $m_L^2=m_D^2+h_3\varphi^2$,
where $h_3$ is the coupling of the $(\phi\phi A_0A_0)$-interaction.
The diagrams with heavy fields in internal lines
do not exist in the final theory, so their effect has to be
produced by different parameters in the action.
One needs to expand the results of
these diagrams in powers of $m/m_D$ to such order that
terms of the form $m_D m$ and $m^2$ are kept.
Constant terms proportional to $m_D^2$ ar neglected.
For the expansion, one writes
\begin{equation}
m_L=m_D+\frac{1}{2}\frac{h_3\varphi^2}{m_D}+\ldots\, .
\end{equation}
There are four types of possible 2-loop
diagrams, and they give~\cite{FKRS1}
\begin{eqnarray}
D_{\rm LS}(m) & = &
\int dp\,dq\frac{1}{(p^2+m_L^2)(q^2+m^2)} \nonumber \\
& = & \bar{I}_3(m_L)\bar{I}_3(m)\Rightarrow
-\frac{\mu^{-2\epsilon}}{4 \pi} m_D \bar{I}_3(m), \label{dls}
\\ & & \nonumber \\
D_{\rm LV}(M) & = &
\int dp\,dq\frac{\delta_{ij}
\Bigl(\delta_{ij}-\frac{q_iq_j}{q^2}\Bigr)}
{(p^2+m_L^2)(q^2+M^2)} \nonumber \\
& = & (2-2\epsilon)\bar{I}_3(m_L)\bar{I}_3(M)\Rightarrow
-\frac{\mu^{-2\epsilon}}{2 \pi} m_D \bar{I}_3(M),
\\ & & \nonumber \\
D_{\rm LLS}(m) & = &
\int dp\,dq\frac{1}
{[p^2+m_L^2][q^2+m_L^2][(p+q)^2+m^2]} \nonumber \\
& = & \bar{H}_3(m_L,m_L,m)\Rightarrow
\frac{\mu^{-4\epsilon}}{16 \pi^2}
\biggl(\frac{1}{4\epsilon}+
\ln\frac{\mu}{2 m_D}+\frac{1}{2}\biggr),
\\ & & \nonumber \\
D_{\rm LLV}(M) & = &
\int dp\,dq\frac{(2p_i+q_i)(2p_j+q_j)
\Bigl(\delta_{ij}-\frac{q_iq_j}{q^2}
\Bigr)}
{[p^2+m_L^2][q^2+M^2][(p+q)^2+m_L^2]} \nonumber \\
& = & (M^2-4m_L^2)\bar{H}_3(m_L,m_L,M)+
2\bar{I}_3(m_L)\bar{I}_3(M)-\bar{I}_3(m_L)\bar{I}_3(m_L) \nonumber \\
& \Rightarrow &
-\frac{\mu^{-2\epsilon}}{\pi} m_D \bar{I}_3(M)\nonumber \\
& & +
\frac{\mu^{-4\epsilon}}{16 \pi^2}
\biggl[
M^2\biggl(\frac{1}{4\epsilon}+
\ln\frac{\mu}{2 m_D}\biggr)-
h_3\varphi^2\biggl(\frac{1}{\epsilon}+
4\ln\frac{\mu}{2 m_D}+1\biggr) \label{dllv}
\biggr].
\end{eqnarray}
Here $\bar{H}_3$ is the $H_3$ of eq.~\nr{l2l}
divided by $T^2$.
The ``linear'' terms proportional to $m_D \bar{I}_3(m)$ in
$D_{\rm LV}$ and $D_{\rm LLV}$ cancel each other, and those in
$D_{\rm LS}$ are cancelled by corrections from $m_D$
to the mass of the scalar field
inside 1-loop diagrams.
This is in complete analogy with the cancellation
of linear terms in the integration over
the superheavy scale by the thermal counterterms.
The actual effect then comes from the diagrams
$D_{\rm LLS}$ and $D_{\rm LLV}$. The $1/\epsilon$-parts
account for the change in the mass counter\-term,
and the rest gives the change
in the renormalized mass parameter.
\section{Dimensional reduction in the Standard Model}
\label{DRinSM}
In this Section we add the correct coefficients,
relevant for the Standard Model,
to the generic results of Sec.~\ref{blocks}.
The coefficients are composed of
combinatorial factors, group theoretic factors from
isospin contractions,
and of coupling constants.
We also explain in detail how the
final 3d theory is constructed.
\subsection{Notation}
We treat the Standard Model
with the Higgs sector
\begin{equation}
{\cal L}_s=(D_{\mu}\Phi)^{\dagger}D_\mu\Phi-\nu^2\Phi^\dagger\Phi+
\lambda\Bigl(\Phi^\dagger\Phi\Bigr)^2
\end{equation}
in the following consistent approximation.
We take $g_Y\neq 0$ only for the top quark, so
that the Yukawa sector is
\begin{equation}
{\cal L}_Y= g_Y \bigl( \bar{q}_L \tilde{\Phi} t_R +
\bar{t}_R \tilde{\Phi}^{\dagger} q_L \bigr).
\end{equation}
Here $\tilde{\Phi}=i\tau_2\Phi^*$, $\tau_2$ is a Pauli matrix,
and $\Phi$ is the Higgs doublet.
The gauge couplings are denoted by~$g$,
$g'$, and $g_S$. We use the formal
power-counting rule $g'^2\sim g^3$, and keep contributions
only below order $g^5$. This allows one to neglect $g'^2$
e.g. inside 1-loop formulas
of vacuum renormalization, so that $m_W=m_Z$ there.
Most of the counterterms of the Standard Model
can be read from eq.~(A11) of~\cite{AE}.
For completeness, let us state the bare parameters of the Higgs sector,
since the terms proportional to $\lambda$ were neglected in~\cite{AE}:
\begin{eqnarray}
\lambda_B & = & \lambda+\frac{\mu^{-2\epsilon}}{(4\pi )^2\epsilon}
\biggl(\frac{9}{16}g^4-\frac{9}{2}\lambda g^2+12\lambda^2
-3g_Y^4+6\lambda g_Y^2\biggr), \\
{\nu}^2_B & = & \nu^2\biggl[1-
\frac{\mu^{-2\epsilon}}{(4\pi )^2\epsilon}\biggl(
\frac{9}{4}g^2-6\lambda-3g_Y^2\biggr)\biggr].
\end{eqnarray}
Here the coupling constants have not been scaled to be dimensionless
in contrast to~\cite{AE}. Note that our
convention for $\lambda$ differs from~\cite{AE}
additionally by the factor~6.
Let us also write down a few useful combinations of the bare
parameters:
\begin{eqnarray}
\nu^2_B\phi_B^2 & = & \nu^2\phi^2
\biggl[1+ \frac{\mu^{-2\epsilon}}{(4\pi )^2\epsilon}
6\lambda
\biggr], \label{dnf} \\
\lambda_B\phi_B^4 & = & \phi^4
\biggl[\lambda+ \frac{\mu^{-2\epsilon}}{(4\pi )^2\epsilon}
\biggl(
\frac{9}{16}g^4+12\lambda^2-3g_Y^4
\biggr)
\biggr], \\
g_B^2\phi_B^2A_B^2 & = & g^2\phi^2A^2
\biggl[1- \frac{\mu^{-2\epsilon}}{(4\pi )^2\epsilon}
\biggl(
\frac{3}{4}g^2+3g_Y^2
\biggr)
\biggr], \\
g_{Y,B}\bar{q}_{L,B}\tilde{\Phi}_Bt_{R,B} & = & g_Y\bar{q}_L
\tilde{\Phi} t_R
\biggl[1- \frac{\mu^{-2\epsilon}}{(4\pi )^2\epsilon}
4 g_S^2
\biggr]. \label{dgY}
\end{eqnarray}
Within our approximation, the Standard
Model contains five running parameters,
$\nu^2(\mu)$, $g^2(\mu)$, $\lambda (\mu)$, $g_Y^2(\mu)$,
and $g_S^2(\mu)$.
The first four run at 1-loop order as
\begin{eqnarray}
\mu\frac{d}{d\mu}\nu^2(\mu) & = & \frac{1}{8\pi^2}
\biggl(-\frac{9}{4}g^2+6\lambda+3g_Y^2\biggr)\nu^2, \label{run1} \\
\mu\frac{d}{d\mu}g^2(\mu) & = & \frac{1}{8\pi^2}
\biggl(\frac{8n_F+N_s-44}{6}\biggr)g^4, \\
\mu\frac{d}{d\mu}\lambda(\mu) & = & \frac{1}{8\pi^2}
\biggl(\frac{9}{16}g^4-\frac{9}{2}\lambda g^2+12\lambda^2-
3g_Y^4+6\lambda g_Y^2\biggr), \\
\mu\frac{d}{d\mu}g_Y^2(\mu) & = & \frac{1}{8\pi^2}
\biggl(\frac{9}{2}g_Y^4-\frac{9}{4}g^2g_Y^2-8g_S^2g_Y^2\biggr).
\label{run4}
\end{eqnarray}
Here $n_F=3$ is the number of families
and $N_s=1$ is the number of Higgs doublets.
The running of the strong
coupling $g_S$ is not explicitly needed for the
present problem, since $g_S$ appears only inside
loop corrections.
The running of $g'^2$ is of higher order
according to our convention.
The fields $\Phi$, $A$ and $\psi$ run as well, and
the formulas can be extracted, e.g., from
eqs.~\nr{dnf}-\nr{dgY},~\nr{run1}-\nr{run4}.
To simplify some of the formulas below, we will
use extensively the notation
\begin{equation}
h\equiv \frac{m_H}{m_W},\quad
t\equiv \frac{m_t}{m_W}, \quad
s\equiv \frac{g_S}{g}, \label{hts}
\end{equation}
where $m_W$, $m_H$ and $m_t$ are the physical
W boson, Higgs particle and top quark masses.
Inside 1-loop formulas one may use the
tree level relations $g^2h^2=8\lambda$ and $g^2t^2=2g_Y^2$.
\subsection{Integration over the superheavy scale}
\label{ioss}
For the SU(2)+Higgs model, the formulas for dimensional
reduction to order~$g^4$ have been given in~\cite{FKRS1}.
We add here the effect of fermions, and correct an error
coming from~\cite{L}.
Due to 3d gauge invariance, the
effective 3d theory is of the form
\begin{eqnarray}
S & = & \int\! d^3x \biggl\{
\frac{1}{4}G^a_{ij}G^a_{ij}+ \frac{1}{4}F_{ij}F_{ij}+
(D_i\Phi)^{\dagger}(D_i\Phi)+
m_3^2\Phi^{\dagger}\Phi+\lambda_3
(\Phi^{\dagger}\Phi)^2 \nonumber \\
& &\hspace*{1.0cm} +\frac{1}{2} (D_iA_0^a)^2+\frac{1}{2}m_D^2A_0^aA_0^a+
\frac{1}{4}\lambda_A(A_0^aA_0^a)^2
+\frac{1}{2} (\partial_iB_0)^2+\frac{1}{2}m_D'^2B_0B_0 \nonumber\\
& &\hspace*{1.0cm}+ h_3\Phi^{\dagger}\Phi A_0^aA_0^a
+ h_3'\Phi^{\dagger}\Phi B_0B_0
-\frac{1}{2}g_3g_3'B_0 \Phi^{\dagger}A_0^a\tau^a\Phi
\,\,\biggr\} ,
\label{action}
\end{eqnarray}
where $G^a_{ij}=\partial_iA_j^a-\partial_jA_i^a+g_3\epsilon^{abc}A^b_iA^c_j$,
$F_{ij}=\partial_iB_j-\partial_jB_i$,
$D_i\Phi=(\partial_i-ig_3\tau^aA^a_i/2+ig_3'B_i/2)\Phi$,
$D_iA_0^a=\partial_iA_0^a+g_3\epsilon^{abc}A_i^bA_0^c$, and
$\Phi=(\phi_3+i\phi_4,\phi_1+i\phi_2)^T/\sqrt{2}$.
The $\tau^a$:s are the Pauli matrices.
The factor $1/T$ multiplying the action has been scaled into
the fields and the coupling constants, so that the fields have
the dimension GeV$^{1/2}$ and the couplings $g_3^2$, $\lambda_3$
have the dimension GeV.
Due to the convention $g'^2\sim g^3$,
we can use $h_3'=g_3'^2/4$, neglect the quartic
coupling of $B_0$-fields,
and use the indicated tree-level values $g_3$, $g_3'$
in the part mixing $A_0$ and $B_0$.
The problem is to calculate the parameters in eq.~\nr{action}
to order~$g^4$ in the coupling constants.
First, let us calculate how the 3d fields are related to the 4d fields.
The momentum-dependent contribution of the superheavy modes
to the two-point scalar correlator is
\begin{equation}
{\cal Z}^{\phi}={\cal Z}^{\phi}_{\rm CT}-
\frac{3}{4}g^2{\cal Z}^{\phi}_{\rm SV}+
\frac{3}{2}g_Y^2{\cal Z}^{\phi}_{\rm FF}=
\frac{k^2}{16\pi^2}(-\frac{9}{4}g^2L_b+3g_Y^2L_f),
\end{equation}
where the ${\cal Z}^{\phi}$'s are from eqs.~\nr{zfct}-\nr{zfff}.
For the spatial and temporal components of the gauge fields one gets
\begin{equation}
{\cal Z}^{A}={\cal Z}^{A}_{\rm CT}-\frac{N_s}{2}g^2{\cal Z}^{A}_{\rm SS}+
2g^2{\cal Z}^{A}_{\eta\eta}
-2n_Fg^2{\cal Z}^{A}_{\rm FF}-g^2{\cal Z}^{A}_{\rm VV},
\end{equation}
where the ${\cal Z}^{A}$'s are from eqs.~\nr{za0ct}-\nr{zaivv}.
This gives
\begin{eqnarray}
{\cal Z}^{A_0} & = &
\frac{g^2k^2}{16\pi^2}\biggl[
\frac{4n_F}{3}(L_f-1)+\frac{N_s}{6}(L_b+2)-\frac{13}{3}L_b +\frac{8}{3}
\biggr],
\\ & & \nonumber \\
{\cal Z}^{A_i} & = &
\biggl(\delta_{ij}-\frac{k_ik_j}{k^2}\biggr)
\frac{g^2k^2}{16\pi^2}\biggl[
\frac{4n_F}{3}L_f+\frac{N_s}{6}L_b-\frac{13}{3}L_b -\frac{2}{3}
\biggr].
\end{eqnarray}
The $\cal Z$'s here correspond just to $\overline{\Pi}'(0)k^2$
in eq.~\nr{pik2}.
Hence the wave functions in the 3d action are related to
the renormalized 4d wave functions in the $\overline{\rm MS}$-scheme by
\begin{eqnarray}
\phi_3^2 & = &
\frac{1}{T}
\phi^2 \biggl\{1+\frac{1}{16\pi^2}\biggl[-\frac{9}{4}g^2L_b+
3g_Y^2L_f\biggr]\biggr\}, \label{phi3} \\
\bigl(A_0^{3d}\bigr)^2 & = &
\frac{1}{T}
\bigl(A_0\bigr)^2\biggl\{1+\frac{g^2}{16\pi^2}\biggl[
\frac{4n_F}{3}(L_f-1)+\frac{N_s}{6}(L_b+2)-\frac{13}{3}L_b +\frac{8}{3}
\biggr]\biggr\}, \label{A03} \\
\bigl(A_i^{3d}\bigr)^2 & = &
\frac{1}{T}
\bigl(A_i\bigr)^2\biggl\{1+\frac{g^2}{16\pi^2}\biggl[
\frac{4n_F}{3}L_f+\frac{N_s}{6}L_b-\frac{13}{3}L_b -\frac{2}{3}
\biggr]\biggr\}. \label{Ai3}
\end{eqnarray}
The factor $1/T$ arises because of the rescaling in eq.~\nr{action}.
The loop corrections to the normalization of $B_0$ are of higher order
according to our convention. When the running of fields in 4d
is taken into account, $\phi_3$, $A_0^{3d}$
and $A_i^{3d}$ are seen to be independent of $\mu$.
The constant~$-2/3$ inside the square brackets
in the formulas for $A_0^{3d}$ and $A_i^{3d}$
in eqs.~\nr{A03}, \nr{Ai3}
is missing in~\cite{FKRS1}, due to
an error in eq.~(6.3) in~\cite{L}. This error propagates
to~$g_3^2$ and~$h_3$; the correct result for $g_3^2$ is
also given in eq.~(6) of~\cite{HL}.
Second, let us calculate the 1-loop corrections to
the coupling constants of the gauge sector.
The couplings $g_3^2$ and $h_3$
can be extracted from the $n\neq 0$ contributions to
the $(\phi\phi A_iA_j)$- and $(\phi\phi A_0A_0)$-correlators
at vanishing external momenta, respectively.
The corrections to the $(\phi\phi B B)$-
and $(\phi\phi B_0A_0)$-vertices are of higher order.
The contributions from the relevant diagrams
are in eqs.~\nr{ga00}-\nr{gaiffff}.
The coefficients are such that
\begin{eqnarray}
{\cal G}^{A} & = &
{\cal G}^{A}_{0}
+{\cal G}^{A}_{\rm CT}
-6\lambda g^2{\cal G}^{A}_{\rm SS}
-g^4{\cal G}^{A}_{\rm SV}
-g^4{\cal G}^{A}_{\rm VV} \nonumber \\
& &
+6\lambda g^2{\cal G}^{A}_{\rm SSS}
+2g^4{\cal G}^{A}_{\rm VVV}
-3g^2g_Y^2{\cal G}^{A}_{\rm FFFF}.
\end{eqnarray}
This gives the effective vertices
\begin{eqnarray}
& & \frac{g^2}{8}\phi_i\phi_iA^a_jA^a_j\biggl[
1+\frac{1}{16\pi^2}
\biggl(\frac{3}{4}g^2L_b+3g_Y^2L_f\biggr)\biggr],\nonumber \\
& & \frac{g^2}{8}\phi_i\phi_iA^a_0A^a_0\biggl[
1+\frac{1}{16\pi^2}
\biggl(\frac{3}{4}g^2L_b+3g_Y^2L_f
+\frac{23}{2}g^2+12\lambda-6g_Y^2
\biggr)
\biggr],
\end{eqnarray}
where the fields are those of the 4d theory.
When the fields are redefined according
to eqs.~\nr{phi3}-\nr{Ai3} and the vertex
is identified with the corresponding vertex in eq.~\nr{action},
one gets the final result for
the coupling constants $g_3^2$ and $h_3$:
\begin{eqnarray}
g_3^2 & = & g^2(\mu)T\biggl[1+\frac{g^2}{16\pi^2}\biggl(
\frac{44-N_s}{6} L_b-
\frac{4n_F}{3} L_f+\frac{2}{3}\biggr)\biggr], \label{g32}\\
h_3 & = & \frac{1}{4}g^2(\mu)T
\biggl[1+ \frac{g^2}{16\pi^2}\biggl(
\frac{44-N_s}{6} L_b-
\frac{4n_F}{3} L_f
+\frac{53}{6}\nonumber \\
& & -\frac{N_s}{3}+\frac{4n_F}{3}
+\frac{3}{2}h^2-3t^2\biggr)\biggr].
\end{eqnarray}
As to the Higgs sector,
the 1-loop unresummed contribution to
the effective potential in Landau gauge is
\begin{equation}
V_1(\varphi)={\cal C}_S(m_1)+3{\cal C}_S(m_2)+
2{\cal C}_V(m_T)+
{\cal C}_V(\sqrt{m_T^2+m_T'^2})+3{\cal C}_F(m_f),
\end{equation}
where the ${\cal C}$'s are from
eqs.~\nr{cs}-\nr{cf}.
The masses appearing in $V_1(\varphi)$ are
\begin{eqnarray}
& & m_1^2 = -\nu^2+3\lambda\varphi^2,\quad
m_2^2=-\nu^2+\lambda\varphi^2,\nonumber \\
& & m_T^2 = \frac{1}{4}g^2\varphi^2,\quad
m_T'^2=\frac{1}{4}g'^2\varphi^2,\quad
m_f^2=\frac{1}{2}g_Y^2\varphi^2. \label{4dms}
\end{eqnarray}
{}From the term quartic in masses in $V_1(\varphi)$,
one gets the $n\neq0$ contribution to the four-point scalar
correlator at vanishing momenta. Redefining the
field $\phi$ according to eq.~\nr{phi3},
the coupling constant $\lambda_3$ is then
\begin{eqnarray}
\lambda_3 & = & \lambda(\mu)T\biggl\{
1-\frac{3}{4}\frac{g^2}{16\pi^2}\biggl[\biggl(
\frac{6}{h^2} -6+2h^2 \biggr)L_b+
\biggl(4 t^2-8\frac{t^4}{h^2}\biggr)L_f-
\frac{4}{h^2} \biggr]\biggr\}.
\end{eqnarray}
The coefficient of $\varphi^2/2$ in $V_1(\varphi)$ gives the 1-loop
result for the scalar mass squared. The result
is the term of order $g^2$ on the first line
of eq.~\nr{m32}.
For the 2-loop contribution to the mass squared $m_3^2$,
one needs the 2-loop effective potential $V_2(\varphi)$.
The diagrams needed for $V_2(\varphi)$ in the Standard Model
are those in Fig.~23 of~\cite{AE},
added by two purely scalar diagrams,
the figure-8 and the sunset. In terms of
eqs.~\nr{fae}-\nr{lae}, the result is
\begin{eqnarray}
V_2(\varphi) & = &
-3\lambda^2\varphi^2\Bigl[{\cal D}_{\rm SSS}(m_1,m_1,m_1)
+{\cal D}_{\rm SSS}(m_1,m_2,m_2)\Bigr]
\nonumber \\ & &
-\frac{3}{8}g^2\Bigl[
{\cal D}_{\rm SSV}(m_1,m_2,m_T)
+{\cal D}_{\rm SSV}(m_2,m_2,m_T)\Bigr]
\nonumber \\ & &
-\frac{3}{64}g^4\varphi^2{\cal D}_{\rm SVV}(m_1,m_T,m_T)
-\frac{1}{2}g^2{\cal D}_{\rm VVV}(m_T,m_T,m_T)
\nonumber \\ & &
-\frac{3}{4}g_Y^2\Bigl[
{\cal D}_{\rm FFS}(m_f,m_f,m_1)
+{\cal D}_{\rm FFS}(m_f,m_f,m_2)
+2{\cal D}_{\rm FFS}(m_f,0,m_2)\Bigr]
\nonumber \\ & &
-\frac{3}{8}g^2\Bigl[
{\cal D}_{\rm FFV}(m_f,m_f,m_T)
+4{\cal D}_{\rm FFV}(m_f,0,m_T)
+(8n_F-5){\cal D}_{\rm FFV}(0,0,m_T)\Bigr]
\nonumber \\ & &
-3g^2{\cal D}_{\eta\eta {\rm V}}(m_T)
-4g_S^2{\cal D}_{\rm FFV}(m_f,m_f,0)
\nonumber \\ & &
-\frac{3}{4}\lambda\Bigl[
{\cal D}_{\rm SS}(m_1,m_1)
+2{\cal D}_{\rm SS}(m_1,m_2)
+5{\cal D}_{\rm SS}(m_2,m_2)\Bigr]
\nonumber \\ & &
-\frac{3}{16}g^2\Bigl[
{\cal D}_{\rm SV}(m_1,m_T)
+3{\cal D}_{\rm SV}(m_2,m_T)\Bigr]
-\frac{3}{4}g^2{\cal D}_{\rm VV}(m_T,m_T)
\nonumber \\ & &
+\frac{1}{2}{\cal D}_{\rm S}(m_1)
+\frac{3}{2}{\cal D}_{\rm S}(m_2)
+\frac{3}{2}{\cal D}_{\rm V}(m_T)
+3{\cal D}_{\rm F}(m_f),
\label{V2sum}
\end{eqnarray}
where the mass counterterms needed
for calculating ${\cal D}_{\rm S}$,
${\cal D}_{\rm V}$, and ${\cal D}_{\rm F}$
include those generated by the shift.
The $\imath_\epsilon$'s from the different
diagrams cancel in the sum. The linear terms
of the form $T^2 I_3(m)$ are cancelled by the thermal
counterterms as explained after eq.~\nr{thctc}.
Apart from vacuum terms,
the terms with $1/\epsilon$'s combine to
$\varphi^2/2$ multiplied by
\begin{equation}
-\frac{T^2}{16\pi^2}\frac{\mu^{-4\epsilon}}{4\epsilon}
\biggl(\frac{81}{16}g^4+9\lambda g^2-12\lambda^2\biggr),\label{premct}
\end{equation}
which is the order $g^4$-result for the mass counterterm of the 3d theory.
However, one can add higher-order corrections to
the mass counterterm by calculating the mass divergence
directly in the 3d theory, obtaining
\begin{equation}
\delta m_3^2=-\frac{1}{16\pi^2}\frac{\mu^{-4\epsilon}}{4\epsilon}
\biggl(\frac{39}{16}g_3^4+12h_3g_3^2-6h_3^2
+9\lambda_3 g_3^2-12\lambda_3^2\biggr).\label{mct}
\end{equation}
This agrees to order $g^4$ with eq.~\nr{premct}
and is the final result for $\delta m_3^2$.
Summing the finite contributions in eq.~\nr{V2sum},
one gets the expression for the renormalized part $m_3^2(\mu)$
of the mass squared. With the shorthand notations
\begin{eqnarray}
\tilde{\nu}^2 & = & \nu^2(\mu)\biggl\{1- \frac{3}{4}\frac{g^2}{16\pi^2}
\biggl[\biggl(
h^2-3\biggr)L_b+2 t^2 L_f\biggr]\biggr\}, \\
\tilde{g}_Y^2 & = & T g_Y^2(\mu)\biggl\{1-\frac{3}{8}
\frac{g^2}{16\pi^2}\biggl[\biggl(6 t^2-
6-\frac{64}{3}s^2\biggr)L_f\nonumber \\
& + & 2+28 \ln{2}-12h^2\ln 2+
8 t^2\ln 2-\frac{64}{9}s^2(4\ln 2-3)\biggr]\biggr\} ,
\end{eqnarray}
the result after redefinition of fields is
\begin{eqnarray}
m_3^2(\mu) & = & -\tilde{\nu}^2
+T\biggl(\frac{1}{2}\lambda_3+\frac{3}{16}g_3^2+\frac{1}{16}g_3'^2+
\frac{1}{4}\tilde{g}_Y^2\biggr)
\nonumber \\
& + &
\frac{T^2}{16\pi^2} \biggl[
g^4\biggl(\frac{137}{96}+\frac{3n_F}{2} \ln{2}+\frac{n_F}{12}\biggr)+
\frac{3}{4}\lambda g^2 \biggr]
\nonumber \\
& + &
\frac{1}{16\pi^2}
\biggl(\frac{39}{16}g_3^4+12h_3g_3^2-6h_3^2+
9\lambda_3g_3^2-12\lambda_3^2\biggr)
\biggl(\ln\frac{3 T}{\mu}+c\biggr).
\label{m32}
\end{eqnarray}
Here, as in eq.~\nr{mct}, we have taken into account higher-order
corrections in the logarithmic term.
The parameters $m_D'^2$, $m_D^2$ and $\lambda_A$ require
the calculation of the effect of $n\neq 0$ modes on the
$(B_0B_0)$-, $(A_0A_0)$- and $(A_0A_0A_0A_0)$-correlators
at vanishing momenta.
Using eqs.~\nr{a2s}-\nr{a2ff}, the
two-point correlator for the U(1)-field $B_0$ is
\begin{eqnarray}
{\cal A}^{(2)}_{\rm U(1)} & = &
g'^2\Bigl[N_s{\cal A}^{(2)}_{\rm S}
-\frac{N_s}{2}{\cal A}^{(2)}_{\rm SS}
-\frac{10n_F}{3}{\cal A}^{(2)}_{\rm FF}\Bigr].\label{a2u1}
\end{eqnarray}
For the SU(2)-field $A_0$, one gets
\begin{eqnarray}
{\cal A}^{(2)}_{\rm SU(2)} & = &
g^2\Bigl[N_s{\cal A}^{(2)}_{\rm S}
+{\cal A}^{(2)}_{\rm V}
-\frac{N_s}{2}{\cal A}^{(2)}_{\rm SS}
-{\cal A}^{(2)}_{\rm VV}
+2{\cal A}^{(2)}_{\eta\eta}
-2n_F{\cal A}^{(2)}_{\rm FF}\Bigr].
\end{eqnarray}
Using eqs.~\nr{a4ss}-\nr{a4ffff}, the
four-point correlator is
\begin{eqnarray}
{\cal A}^{(4)}_{\rm SU(2)} & = &g^4\Bigl[
-\frac{N_s}{4}{\cal A}^{(4)}_{\rm SS}
+\frac{N_s}{2}{\cal A}^{(4)}_{\rm SSS}
-\frac{N_s}{8}{\cal A}^{(4)}_{\rm SSSS}
-\frac{1}{2}{\cal A}^{(4)}_{\rm VV} \nonumber \\
& &
+2{\cal A}^{(4)}_{\rm VVV}
-{\cal A}^{(4)}_{\rm VVVV}
+2{\cal A}^{(4)}_{\eta\eta\eta\eta}
+\frac{n_F}{2}{\cal A}^{(4)}_{\rm FFFF}\Bigr].\label{a4su2}
\end{eqnarray}
Since there is no tree-level term corresponding to
the correlators in eqs.~\nr{a2u1}-\nr{a4su2},
the redefinition of fields in eq.~\nr{A03} produces
terms of higher order. The final results
can then be read directly from eqs.~\nr{a2u1}-\nr{a4su2}:
\begin{eqnarray}
m_D'^2 & = & \biggl(\frac{N_s}{6}+\frac{5n_F}{9}\biggr) g'^2 T^2, \\
m_D^2 & = & \biggl(\frac{2}{3}+\frac{N_s}{6}+\frac{n_F}{3}\biggr) g^2 T^2, \\
\lambda_A & = & T\frac{g^4}{16\pi^2} \frac{16+N_s-4 n_F}{3}.
\end{eqnarray}
In principle, the mass $m_D^2$ should be determined to
order $g^4$ to be compatible with the accuracy of
vacuum renormalization. We have, however, not made this calculation,
since the effect of $g^4$-corrections to $m_D^2$
contributes in higher order
than $g^4$ to the Higgs field effective potential $V(\varphi)$,
which drives the EW phase transition.
Using eqs.~\nr{run1}-\nr{run4}, one
sees that the quantities $g_3^2$, $h_3$,
$\lambda_3$, $\tilde{\nu}^2$ and $\tilde{g}_Y^2$
are independent of $\mu$ to the
order they are presented above. In other words,
when the running parameters
$g^2(\mu)$, $\nu^2(\mu)$, $\lambda(\mu)$
and $g_Y^2(\mu)$ are expressed in terms of physical
parameters in Sec.~\ref{MSbar-Ph}, the $\mu$-dependence
cancels in the 3d parameters.
The $\mu$-dependence of $\lambda_A$, $m_D^2$ is of higher order,
as well; actually $m_D^2$ runs only at order $g^6$~\cite{FKRS1}.
Note also that the $\mu$ in $m_3^2(\mu)$ is independent of the
$\mu$ used in the construction of the 3d theory (although the
notation is the same), since the bare mass parameter
$m_3^2(\mu)+\delta m_3^2$,
being the sum of eqs.~\nr{mct} and \nr{m32}, is independent of $\mu$.
\subsection{Integration over the heavy scale}
\label{iohs}
The action in eq.~\nr{action} can be further simplified
by integrating out the $A_0$- and $B_0$-fields.
The masses of these fields are of the order $gT$ and
$g'T\sim g^{3/2}T$, respectively.
The resulting action is of the form in eq.~\nr{action}
with the $A_0$- and $B_0$-fields left out and the parameters modified.
There are no infrared divergences related to this integration, either,
since the $A_0$- and $B_0$-fields are massive.
We denote the new parameters by a bar.
For the SU(2)-Higgs model, the relations of
the old and the barred parameters have been given
in~\cite{FKRS1}.
The calculation of the barred parameters proceeds
in complete analogy with dimensional reduction.
At 1-loop level, there is no momentum-dependent
correction from the heavy $A_0$- and $B_0$-fields
to the $\phi_3$-correlator,
so that $\bar{\phi}_3=\phi_3$. The momentum-dependent
correction to the $A_i$-correlator is
\begin{equation}
-g_3^2{\cal Z}^{A_i}_{\rm LL},
\end{equation}
where ${\cal Z}^{A_i}_{\rm LL}$ is from eq.~\nr{zaill}, so that
\begin{equation}
\bigl(\bar{A}_i\bigr)^2 = \bigl(A_i^{3d}\bigr)^2
\biggl(1+\frac{g_3^2}{24\pi m_D}\biggr).
\end{equation}
The field $B_0$ does not get normalized, since $B_0$ and $B_i$
do not interact.
To find $\bar{g}_3^2$, one can evaluate the two diagrams
with $A_0$ in the loop contributing to the
$(\phi_3)^2 \bigl(A^{3d}_{i}\bigr)^2$-vertex.
In terms of ${\cal G}_{\rm LL}$
and ${\cal G}_{\rm LLL}$ in eqs.~\nr{gll}-\nr{glll},
the result is
\begin{equation}
g_3^2\delta_{ij}-8h_3g_3^2{\cal G}_{\rm LL}+
8h_3g_3^2{\cal G}_{\rm LLL}=
g_3^2\delta_{ij}.
\end{equation}
Hence
\begin{equation}
\bar{g}_3^2\bar{\phi}_3^2\bar{A}_i^2=
g_3^2\phi_3^2\bigl(A^{3d}_{i}\bigr)^2,
\end{equation}
which gives
\begin{equation}
\bar{g}_3^2=g_3^2\biggl(
1-\frac{g_3^2}{24\pi m_D}\biggr).
\end{equation}
The coupling $g_3'^2$ does not get normalized,
since $B_0$ and $B_i$ do not interact.
The 1-loop corrections to the scalar coupling constant are
\begin{equation}
-3h_3^2{\cal B}(0;m_D,m_D)
-\frac{1}{8}g_3^2g_3'^2{\cal B}(0;m_D,m_D')
-\frac{1}{16}g_3'^4{\cal B}(0;m_D',m_D').
\end{equation}
This gives
\begin{equation}
\bar{\lambda}_3=\lambda_3-\frac{1}{8\pi}\biggl(
3\frac{h_3^2}{m_D}+
\frac{g_3'^4}{16 m_D'}+\frac{g_3'^2 g_3^2}{4(m_D+m_D')}
\biggr).
\end{equation}
The 1-loop corrections to the scalar mass parameter are
\begin{equation}
3h_3\bar{I}_3(m_D)+\frac{1}{4}g_3'^2\bar{I}_3(m_D'),
\end{equation}
giving the first line in eq.~\nr{bm32}.
To calculate the 2-loop corrections, one needs the
effective potential. The 2-loop contribution from
the heavy scale to the effective potential is
\begin{eqnarray}
V_2^{\rm heavy}(\varphi) & = &
\frac{3}{2}h_3\Bigl[
D_{\rm LS}(m_1)+3D_{\rm LS}(m_2)\Bigr]
+3g_3^2D_{\rm LV}(m_T) \nonumber \\
& & -3h_3^2\varphi^2D_{\rm LLS}(m_1)
-\frac{3}{2}g_3^2 D_{\rm LLV}(m_T),
\end{eqnarray}
where the $D$'s are from eqs.~\nr{dls}-\nr{dllv}
and we used $g'^2\sim g^3$.
The $1/\epsilon$-parts modify the mass counterterm
of eq.~\nr{mct} to become
\begin{equation}
-\frac{1}{16\pi^2}\frac{\mu^{-4\epsilon}}{4\epsilon}
\biggl(\frac{51}{16}g_3^4
+9\lambda_3 g_3^2-12\lambda_3^2\biggr).
\end{equation}
However, one can again include higher order corrections
by calculating the counterterm directly in the final theory,
getting
\begin{equation}
\delta\bar{m}_3^2=-\frac{1}{16\pi^2}\frac{\mu^{-4\epsilon}}{4\epsilon}
\biggl(\frac{51}{16}\bar{g}_3^4
+9\bar{\lambda}_3 \bar{g}_3^2-12\bar{\lambda}_3^2\biggr).\label{bcmt}
\end{equation}
Finally, the renormalized
mass parameter $\bar{m}_3^2(\mu)$ is
\begin{eqnarray}
\bar{m}_3^2(\mu) & = &m_3^2(\mu)-\frac{1}{4\pi}\biggl(3 h_3 m_D +
\frac{1}{4}g_3'^2m_D'\biggr) \nonumber \\
& &\hspace*{0.5cm} +\frac{1}{16\pi^2}\biggl[
\biggl(-\frac{3}{4}g_3^4+12h_3g_3^2-6h_3^2\biggr)
\ln\frac{\mu}{2m_D}+3h_3g_3^2-3h_3^2\biggr] \nonumber \\
& & \nonumber \\
& = & -\tilde{\nu}^2
+T\biggl(\frac{1}{2}\lambda_3+\frac{3}{16}g_3^2+\frac{1}{16}g_3'^2+
\frac{1}{4}\tilde{g}_Y^2\biggr)
-\frac{1}{4\pi}\biggl(3 h_3 m_D +
\frac{1}{4}g_3'^2m_D'\biggr) \nonumber \\
& & \hspace*{0.5cm}
+\frac{T^2}{16\pi^2} \biggl[
g^4\biggl(\frac{137}{96}+\frac{3n_F}{2} \ln{2}+\frac{n_F}{12}\biggr)+
\frac{3}{4}\lambda g^2\biggr]
\nonumber \\
& & \hspace*{0.5cm}
+\frac{1}{16\pi^2}
\biggl[
\biggl(-\frac{3}{4}g_3^4+12h_3g_3^2-6h_3^2
\biggr)\biggl(\ln\frac{3 T}{2 m_D}+c\biggr)
+3h_3g_3^2-3h_3^2
\biggr]
\nonumber \\
& & \hspace*{0.5cm}
+\frac{1}{16\pi^2}
\biggl(\frac{51}{16}\bar{g}_3^4+
9\bar{\lambda}_3\bar{g}_3^2-12\bar{\lambda}_3^2\biggr)
\biggl(\ln\frac{3 T}{\mu}+c\biggr),
\label{bm32}
\end{eqnarray}
where we used eq.~\nr{m32}
and included higher order corrections in the logarithmic
term on the last line.
Eq.~\nr{bm32} completes the evaluation of the couplings of
the 3d SU(2)$\times$U(1)+Higgs theory.
\section{Relation of the $\overline{\rm MS}$-scheme and Physics in
the Standard Model}
\label{MSbar-Ph}
In Sec.~\ref{DRinSM}, we have given the relations of the running parameters
in the $\overline{\mbox{\rm MS}}$-scheme to the parameters of the
effective 3d theory with accuracy $g^4$. For this accuracy to be
meaningful, the running parameters in the $\overline{\rm MS}$-scheme
should be expressed with accuracy $g^4$ in terms of true physical
parameters, like pole masses. We give these relations in the
present Section.
The accuracy $g^4$ requires 1-loop renormalization
of the vacuum theory. The 1-loop renormalization of the
Standard Model is, of course, a well studied subject.
Usually, however, one does not use the $\overline{\mbox{\rm MS}}$-scheme
we have employed above,
but the so called on-shell scheme~\cite{S1,S2,H1,H2}.
In the on-shell scheme, the divergences appearing in the
loop integrals are handled with dimensional regularization
as in the $\overline{\rm MS}$-scheme, but the finite parts
of the counterterms are chosen differently. Indeed,
the renormalized parameters are chosen to be physical
and independent of $\mu$. For instance, the renormalized
mass squared of the scalar field is $\nu^2_{\rm os}=m_H^2/2$,
and the gauge and scalar couplings are given by
\begin{equation}
g^2_{\rm os}=\frac{e^2_{\rm os}}{s^2_W}, \quad
g'^2_{\rm os}=\frac{e^2_{\rm os}}{c^2_W}, \quad
\lambda_{\rm os}=\frac{g^2_{\rm os}}{8}\frac{m_H^2}{m_W^2}.\label{os}
\end{equation}
Here $m_H$ and $m_W$ are the physical pole masses of the
Higgs particle and the W boson, and
\begin{equation}
e^2_{\rm os}\equiv 4 \pi\alpha,\quad
c_W\equiv\frac{m_W}{m_Z},\quad
s_W\equiv\sqrt{1-\frac{m_W^2}{m_Z^2}},\label{notation}
\end{equation}
where $\alpha$ is the electromagnetic fine structure constant
defined in the Thomson limit, and $m_Z$ is the physical pole mass
of the Z boson. The parameters $c_W$ and $s_W$ are just
shorthand notations without any higher order corrections.
To convert results from the on-shell scheme
to the $\overline{\rm MS}$-scheme, let us note that since
all physical quantities derived in the two schemes
are exactly the same, the bare Lagrangians, including
the counterterms, must be the same.
In particular, all the bare parameters of the theories
must be the same (often the wave function renormalization
factors are not even needed, see e.g.~\cite{S1}).
For the gauge coupling this means
\begin{equation}
g_B^2=g^2(\mu)+\delta g^2(\mu)=g^2_{\rm os}+\delta g^2_{\rm os},
\end{equation}
that is,
\begin{equation}
g^2(\mu)=g^2_{\rm os}\biggl(1+\frac{\delta g^2_{\rm os}-
\delta g^2(\mu)}{g^2_{\rm os}}\biggr)
\label{g2mu}.
\end{equation}
The counterterm $\delta g^2(\mu)$ of the $\overline{\rm MS}$-scheme
cancels the $1/\epsilon$-parts in $\delta g^2_{\rm os}$, so that
eq.~\nr{g2mu} is finite. In this way, the running parameters
in the $\overline{\rm MS}$-scheme can be expressed
in terms of physical parameters. We shall work out the
explicit expressions in some detail below; the reader
only interested in the final numerical results
for the 3d parameters in terms of the
physical 4d parameters may turn to Sec.~\ref{results}.
\subsection{The parameters $g^2(\mu)$ and $g'^2$ in terms of
physical parameters}
\label{ggprim}
The expression for $\delta g^2_{\rm os}$, determining
$g^2(\mu)$ through eq.~\nr{g2mu}, can in principle
be read directly for example from eq.~(28.a) in~\cite{S1}.
With our sign conventions, the equation reads
\begin{equation}
\frac{\delta g^2_{\rm os}}{g^2_{\rm os}}=
2\frac{\delta e_{\rm os}}{e_{\rm os}}-
\frac{c_W^2}{s_W^2}\biggl[\frac{\tilde{\Pi}_Z(-m_Z^2)}{m_Z^2}-
\frac{\tilde{\Pi}_W(-m_W^2)}{m_W^2}\biggr]. \label{dg2}
\end{equation}
Here $\tilde{\Pi}$ means the unrenormalized but regularized self-energy.
This equation is gauge-independent, since $\delta e_{\rm os}$
and the self-energies at the pole are~\cite{S1}.
However, a reliable estimate of eq.~\nr{dg2}
cannot be given purely perturbatively, since
the expression for $\delta e_{\rm os}$ contains
the photon self-energy $\tilde{\Pi}_{\gamma}(k^2)/k^2$ evaluated
at vanishing momentum $k^2$. Indeed, $\tilde{\Pi}_{\gamma}(k^2)/k^2$
includes logarithms of all the small lepton and quark masses,
as is seen e.g. in eq.~(5.40) of~\cite{H1}. This
indicates that strong interactions are important for $\delta e_{\rm os}$.
There exists a standard technique of expressing
the hadronic contribution to $\delta e_{\rm os}$
in terms of a dispersion relation, and hence
$\delta e_{\rm os}$ is known with very good accuracy~\cite{H2,C}
in spite of strong interactions. We find it convenient, though,
to write the expression for $\delta g^2_{\rm os}$ in
a form somewhat different from eq.~\nr{dg2}, so that $\delta e_{\rm os}$ is
not directly visible. Such a way is expressing
$\delta g^2_{\rm os}$ in terms of the Fermi constant $G_\mu$.
The Fermi constant $G_\mu$ is defined~\cite{S1,S2,C}
in terms of the muon lifetime by
\begin{equation}
\frac{1}{\tau_\mu}=\frac{G_\mu^2m_\mu^5}{192 \pi^3}
\biggl(1-8\frac{m_e^2}{m_\mu^2}\biggr)
\biggl[1+\frac{\alpha}{2\pi}
\biggl(1+\frac{2\alpha}{3\pi}\ln\frac{m_\mu}{m_e}\biggl)
\biggl(\frac{25}{4}-\pi^2\biggr)\biggr].\label{Gmu}
\end{equation}
The $\alpha$-corrections here account for the QED-corrections to
muon decay in the local Fermi-model. On the other hand,
calculating the muon lifetime in the Standard Model, one
gets a prediction for the $G_\mu$ of eq.~\nr{Gmu}~\cite{S1,S2,H1,H2}:
\begin{equation}
\frac{G_\mu}{\sqrt{2}}=\frac{g^2_{\rm os}}{8 m_W^2}\frac{1}{1-\Delta r}.
\label{Gu}
\end{equation}
Here $\Delta r$ depends on the parameters of the Standard Model.
Using eqs.~(28.a),~(34.b) of~\cite{S1},
the 1-loop expression for $\Delta r$ in the on-shell scheme
in the Feynman-$R_\xi$-gauge can be written as
\begin{eqnarray}
\Delta r & = & \frac{{\mathop {\rm Re}}
\Bigl[\tilde{\Pi}_W(0)-\tilde{\Pi}_W(-m_W^2)\Bigr]}{m_W^2}+
\frac{\delta g^2_{\rm os}}{g^2_{\rm os}} \nonumber \\
& + & \frac{g^2_{\rm os}}{16 \pi^2}
\biggl[
4\biggl(\frac{1}{\epsilon}+\ln\frac{\mu^2}{m_W^2}\biggr)
+6+\frac{7-4 s_W^2}{2s_W^2}\ln c_W^2\biggr].
\label{Dr}
\end{eqnarray}
This equation can also be extracted from~\cite{H2} as a combination of
eqs.~(3.8), (3.16), (3.17), (4.18), (A.1), (A.2) and (B.3).
Let us note that the value of eq.~\nr{Dr}
is finite and gauge-independent, since~$\Delta r$
is a physical observable.
In addition, $\delta g^2_{\rm os}/g^2_{\rm os}$ and
$\tilde{\Pi}_W(-m_W^2)$ are gauge-independent. However,
$\tilde{\Pi}_W(0)$ is not gauge-independent, and eq.~\nr{Dr}
as a whole holds only in the Feynman-$R_\xi$ gauge.
Usually eq.~\nr{Dr}, with $\delta g^2_{\rm os}$
plugged in from eq.~\nr{dg2} and $g^2_{\rm os}$ from
eqs.~\nr{os},~\nr{notation},
is used in determining $m_W$ from eq.~\nr{Gu}
in terms of the very precisely known
parameters $G_\mu$ and $m_Z$,
and the masses $m_t$ and $m_H$~\cite{H2,C}.
For fixed $m_t$ and $m_H$, the estimated uncertainty
in the value of $m_W$ determined this way is only of
the order of $0.01$~GeV~\cite{H2,C,SIII,K}.
This is much smaller than the present experimental uncertainty
in the W mass, $m_W=80.22\pm 0.18$~GeV~\cite{C}.
In other words, with the present accuracy in the determination
of $m_W$, one should replace $m_W$ as an input parameter
with the hadronic contribution to $\delta e_{\rm os}$.
The $m_W$ obtained this way is shown in Table~\ref{WHrel}
as a function of $m_t$ and $m_H$ for $\alpha_S=0.125$.
\begin{table}[htbp]
\centering
\begin{minipage}[t]{13cm}
\caption[a]{\protect
The mass $m_W$ as a function of $m_t$ and $m_H$ according to~\cite{T}.
The uncertainty in $m_W$ is about $0.01$~GeV for fixed
$m_t$, $m_H$~\cite{H2,C,SIII,K}.}
\vspace*{4mm}
\end{minipage}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|} \hline
$m_t\backslash m_H$ &
35 & 50 & 60 & 70 & 80 & 90 & 100 & 200 & 300 \\ \hline
165 &
80.38 & 80.36 & 80.35 & 80.34 & 80.34 & 80.33 & 80.32 & 80.28 & 80.25 \\ \hline
175 &
80.44 & 80.42 & 80.41 & 80.41 & 80.40 & 80.39 & 80.39 & 80.34 & 80.31 \\ \hline
185 &
80.50 & 80.49 & 80.48 & 80.47 & 80.46 & 80.46 & 80.45 & 80.40 & 80.37 \\ \hline
\end{tabular}
\label{WHrel}
\end{table}
When $m_W$ is fixed, either from Table~1
or in the future from experiment,
$\Delta r$ is known from eq.~\nr{Gu},
and $\delta g^2_{\rm os}$
can be solved from eq.~\nr{Dr} in a simple form.
Using eq.~\nr{g2mu}, one then gets
\begin{eqnarray}
g^2(\mu) & = & g^2_0\biggl\{1+
\frac{g^2_0}{16 \pi^2}
\biggl[\biggl(\frac{4 n_F}{3}-\frac{43}{6}\biggr)
\ln\frac{\mu^2}{m_W^2} \nonumber\\
& - & \frac{33}{4} F(m_W;m_W,m_W)+
\frac{1}{12}(h^4-4h^2+12)F(m_W;m_W,m_H) \nonumber\\
& - &
\frac{1}{2}(t^4+t^2-2)F(m_W;m_t,0)
-2\ln{t}-\frac{h^2}{24} +\frac{t^2}{4} +\frac{20 n_F}{9}-
\frac{257}{72}\biggr] \biggr\}.\label{fineq}
\end{eqnarray}
Here we used $g'^2\sim g^3$, and
defined $g_0^2=4\sqrt{2}G_\mu m_W^2=g^2_{\rm os}/(1-\Delta r)$.
The reason for using $g_0^2$ as the tree-level value instead of
$g^2_{\rm os}$ is that for $g^2_{\rm os}$ there would be a
rather large correction $\Delta r$
in the 1-loop formula, indicating bad convergence.
For instance, for $m_W=80.22$~GeV, $\Delta r =0.045$.
The physical reason~\cite{S2,H2,C} for
the large correction is that $g^2_{\rm os}$ is defined
in terms of the fine structure constant $\alpha$ measured at vanishing
momentum scale, whereas the momentum scale of weak interactions
is~$m_W^2$. With~$g_0^2$,
the 1-loop correction is extremely small for $\mu\sim m_W$.
The function $F(k;m_1,m_2)$ in eq.~\nr{fineq}
has been defined in~\cite{H1},
and its explicit form is given in eq.~\nr{Fkm}.
Note that from eqs.~\nr{Dr} and~\nr{fineq} one can see
that there are no dangerous logarithms in $\tilde{\Pi}_W(0)/m_W^2$
in contrast to $\tilde{\Pi}_\gamma(0)/k^2$, since
any such logarithms are suppressed by $m_f^2/m_W^2$.
Eq.~\nr{fineq} is the final result for $g^2(\mu)$
in terms of physical parameters.
In analogy with the definition for $g_0^2$, we define the U(1)
coupling to be
\begin{equation}
g'^2=g_0^2\frac{s_W^2}{c_W^2}=g_0^2\frac{m_Z^2-m_W^2}{m_W^2} \label{gprim}
\end{equation}
instead of $g'^2_{\rm os}$. The value of $g'^2$ is
larger than the value of $g'^2_{\rm os}$, corresponding
again roughly to $\alpha_{\rm EM}$ running from the Thomson limit $q^2=0$
to the electroweak scale.
\subsection{The parameters $\nu^2(\mu)$, $\lambda(\mu)$,
and $g_Y^2(\mu)$ in terms of physical parameters}
\label{nulgY}
The parameters $\nu^2(\mu)$, $\lambda(\mu)$,
and $g_Y^2(\mu)$ could be determined from the precision
calculations in the on-shell scheme just as $g^2(\mu)$.
For illustration, however, we will calculate the 1-loop corrections
to the tree-level values
\begin{equation}
\nu^2=\frac{1}{2}m_H^2,\quad
\lambda=\frac{1}{\sqrt{2}}G_\mu m_H^2=\frac{g_0^2}{8}
\frac{m_H^2}{m_W^2},\quad
g_Y^2=2\sqrt{2}G_\mu m_t^2=\frac{g_0^2}{2}\frac{m_t^2}{m_W^2} \label{0l}
\end{equation}
in some more detail.
To calculate $\nu^2(\mu)$, $\lambda(\mu)$, and $g_Y^2(\mu)$
at 1-loop level, one has to calculate the 1-loop corrections
to the propagators of the Higgs particle, $W$-boson and
top quark in the broken phase, and extract from these
the pole masses. The diagrams needed are shown in Fig.~\ref{pi}.
To go to the broken phase, the Higgs field $\phi_1$
is shifted to the classical minimum $\varphi$,
where $\varphi^2=\nu^2/\lambda$.
The masses appearing in the Feynman rules are denoted
by $m_1^2=2\nu^2$ for the Higgs field,
$m_T^2=g^2\nu^2/4\lambda$ for the $W$-boson,
and $m_f^2=g_Y^2\nu^2/2\lambda$ for the top quark.
The radiatively corrected 1-loop propagators are then of the form
\begin{eqnarray}
\langle\phi_1(-p)\phi_1(p)\rangle & = & \frac{1}{p^2+m_1^2-\Pi_H(p^2)},
\nonumber \\
\langle A^a_\mu(-p)A^b_\nu(p)\rangle & = &
\delta^{ab}\frac{\delta_{\mu\nu}-p_\mu p_\nu/p^2}
{p^2+m_T^2-\Pi_W(p^2)}
+ {\rm \quad longitudinal\; part} \label{prop} \\
\langle\Psi_\alpha(p)\overline{\Psi}_\beta(p)\rangle & = &
\biggl[\frac{1}{i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}+m_f+i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\Sigma_v(p^2)+
i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_5\Sigma_a(p^2)+
m_f\Sigma_s(p^2)}\biggr]_{\alpha\beta}. \nonumber
\end{eqnarray}
To give the results for the radiatively corrected propagators,
we use the function $F(k;m_1,m_2)$ defined in~\cite{H1}. For
$|m_1-m_2|<k<m_1+m_2$, $F(k;m_1,m_2)$ is
\begin{eqnarray}
F(k;m_1,m_2) & = & 1-\frac{m_1^2-m_2^2}{k^2}\ln\frac{m_1}{m_2}+
\frac{m_1^2+m_2^2}{m_1^2-m_2^2}\ln\frac{m_1}{m_2} \label{Fkm} \\
& - & \frac{2}{k^2}\sqrt{(m_1+m_2)^2-
k^2}\sqrt{k^2-(m_1-m_2)^2}\arctan\frac{\sqrt{k^2-(m_1-m_2)^2}}
{\sqrt{(m_1+m_2)^2-k^2}},\nonumber
\end{eqnarray}
and has the special values
\begin{eqnarray}
F(m_1;m_1,m_2) & = & 1-r^2\frac{3-r^2}{1-r^2}\ln{r}-
2 r \sqrt{4-r^2}\arctan\frac{\sqrt{2-r}}{\sqrt{2+r}}, \nonumber \\
F(m_1;m_2,m_2) & = & 2-2 \sqrt{4 r^2-1}\arctan\frac{1}{\sqrt{4 r^2-
1}}, \nonumber \\
F(m;m,m) & = & 2-\frac{\pi}{\sqrt{3}},
\end{eqnarray}
where $r=m_2/m_1$. We also need the analytical continuation
\begin{equation}
F(m_1;m_2,0) = 1+(r^2-1)\ln\biggl(1-\frac{1}{r^2}\biggr).
\end{equation}
For $m_H<2m_W$, the only formula needed in the region
where it develops an imaginary part, is $F(m_1;m_2,0)$.
With the help of $F(k;m_1,m_2)$, one can write down the special values
\begin{eqnarray}
\Pi_H(-m_H^2) & = & \!\!
\frac{3}{8}\frac{g^2}{16 \pi^2}m_H^2
\biggl[2(h^2+2t^2-3)\ln\frac{\mu^2}{m_W^2} +
3 h^2 F(m_H;m_H,m_H) \nonumber \\
& + & \!\!\frac{h^4-4h^2+12}{h^2}F(m_H;m_W,m_W)-
4 t^2 \frac{4t^2-h^2}{h^2}F(m_H;m_t,m_t)\nonumber \\
& - & \!\!2 h^2 \ln{h}-8 t^2 \ln{t}-
2 h^2-2-12\frac{1}{h^2}+16 \frac{t^4}{h^2}\biggr],\label{PH}\\
\Pi_W(-m_W^2) & = & \frac{3}{8}\frac{g^2}{16 \pi^2}m_W^2
\biggl[ 2\Bigl( \frac{16 n_F-59}{9}-
h^2-2 t^2-6\frac{1}{h^2}+
8 \frac{t^4}{h^2}\Bigl)\ln\frac{\mu^2}{m_W^2} \nonumber\\
& - & 22 F(m_W;m_W,m_W)+
\frac{2}{9}(h^4-4h^2+12)F(m_W;m_W,m_H) \nonumber\\
& - &
\frac{4}{3}(t^4+t^2-2)F(m_W;m_t,0) \nonumber\\
& + & 4 h^2 \frac{h^2-2}{h^2-1} \ln{h}+
\frac{8}{3}\Bigl(3t^2-2-12\frac{t^4}{h^2}\Bigr)\ln{t}-
\frac{22}{9} h^2-\frac{4}{h^2}-\frac{4}{3} t^2 \nonumber\\
& + & 16 \frac{t^4}{h^2}+
\frac{4}{27}(40 n_F-17)+
\frac{8}{3}\Bigl(1-\frac{4}{3} n_F\Bigr)\ln(-1-i\epsilon)\biggr], \label{PW}\\
\Sigma_v(-m_t^2)-\Sigma_s(-m_t^2) & = &
\frac{3}{16}\frac{g^2}{16 \pi^2}
\biggl[2\Bigl(t^2-h^2-6\frac{1}{h^2}+
8\frac{t^4}{h^2}-\frac{32}{3}s^2
\Bigr)\ln\frac{\mu^2}{m_W^2}\nonumber\\
& + &\frac{2}{3}(4t^2-h^2)F(m_t;m_t,m_H)
+\frac{2}{3}\Bigl(1-\frac{1}{t^2}\Bigr)F(m_t;m_t,m_W)\nonumber \\
& + & \frac{2}{3}\Bigl(t^2+1-\frac{2}{t^2}\Bigr)F(m_t;m_W,0) \nonumber\\
& + & 4 h^2 \ln{h}-32 \frac{t^4}{h^2} \ln{t}-
\frac{4}{3}t^2\frac{2t^2+h^2}{t^2-h^2}\ln\frac{t}{h}+
\frac{128}{3}s^2\ln t \nonumber\\
& - & 2+2t^2-2h^2-\frac{4}{h^2}+16\frac{t^4}{h^2}
-\frac{256}{9}s^2\biggr]. \label{PF}
\end{eqnarray}
Here we again used notation from eq.~\nr{hts}.
The expressions~\nr{PH}-\nr{PF} are gauge-independent,
and are the only values of $\Pi_H$, $\Pi_W$ and the $\Sigma$'s
that will be needed here. The value of $\Pi_W(-m_W^2)$ can
be extracted from~\cite{H1}, and that
of $\Pi_H(-m_H^2)$ from~\cite{CEQR}. As a check, we have explicitly
verified the gauge-independence of $\Pi_W(-m_W^2)$.
Numerically the dominant terms in eqs.~\nr{PH}-\nr{PF}
are the fermionic contributions $16t^4/h^2$ and
$-32(t^4/h^2)\ln(\mu/m_t)$. These terms come from
the fermionic tadpoles, and from the fermionic loops in the
Higgs and $W$-boson correlators. The fermionic terms
are large because $m_t/m_W$ is large,
and hence the gauge coupling is large: $g_Y\approx 1.0$.
Consequently, higher order fermionic corrections
are rather important.
With eqs.~\nr{PH}-\nr{PF} and~\nr{fineq}, one can
express $\nu^2(\mu)$, $\lambda (\mu)$ and $g_Y^2(\mu)$
in terms of physical parameters. To do so, the physical
pole masses have to be extracted from eqs.~\nr{prop}.
For the Higgs particle and $W$-boson, this is
straightforward. For the top quark, the 1-loop
equation for the physical mass $m_t$ is
\begin{equation}
\overline{u}(p)
\bigl[i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}+m_f+i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\Sigma_v(-m_t^2)+
i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}\gamma_5\Sigma_a(-m_t^2)+
m_f\Sigma_s(-m_t^2)\bigr]u(p)=0. \label{teq}
\end{equation}
Here $u(p)$ is an asymptotic spinor
satisfying $(i\@ifnextchar[{\@slash}{\@slash[\z@]}{p}+m_t)u(p)=0$.
In eq.~\nr{teq}, the factor $\Sigma_a$ multiplying $\gamma_5$
does not affect the physical mass, since
$\bar{u}(p)\gamma_5 u(p)=0$. The factor $\Sigma_a$ would have an effect
if the top mass were determined from the requirement
that the determinant of the inverse propagator vanishes, in
which case $\Sigma_a$ produces an
unphysical imaginary part to the self-energy even for $m_t<m_W$.
Physically, the reason why the top quark can be considered
an asymptotic state, is that the time scale~$\tau\sim m_W^4/m_t^5$
of weak interactions is much smaller
than the scale $1$~fm of strong interactions.
Evaluating the expressions for the
Higgs particle, W boson, and top quark masses,
and expressing $m_1$, $m_T$ and $m_f$ in terms of the coupling constants,
one can then solve for the parameters $\nu^2(\mu)$,
$\lambda(\mu)$ and $g_Y^2(\mu)$. The results are
\begin{eqnarray}
\nu^2(\mu) & = &\frac{m_H^2}{2}\,{\mathop{\rm Re}}
\biggl[1+\frac{\Pi_H(-m_H^2)}{m_H^2}\biggr], \nonumber \\
\lambda (\mu) & = & \frac{g_0^2}{8}\frac{m_H^2}{m_W^2}
\,{\mathop{\rm Re}} \biggl[1-
\frac{\Pi_W(-m_W^2)}{m_W^2}+\frac{\delta g^2(\mu)}{g_0^2}
+\frac{\Pi_H(-m_H^2)}{m_H^2}\biggr], \label{physMS} \\
g_Y^2(\mu) & = & \frac{g_0^2}{2}\frac{m_t^2}{m_W^2}
\,{\mathop{\rm Re}} \biggl[
1-\frac{\Pi_W(-m_W^2)}{m_W^2}+\frac{\delta g^2(\mu)}{g_0^2}
+2\Sigma_v(-m_t^2)-2\Sigma_s(-m_t^2)\biggr].
\nonumber
\end{eqnarray}
Here $\delta g^2(\mu)$ is the renormalized
1-loop correction in eq.~\nr{fineq}.
The $\mu$-dependences in eqs.~\nr{physMS}
naturally reproduce those in eqs.~\nr{run1}-\nr{run4}.
We will not write down explicitly the expressions
in eq.~\nr{physMS}, since
we did not find any significant simplification
in the final result, apart from the $\mu$-dependent terms.
Together with eqs.~\nr{fineq} and~\nr{gprim} and
the value $\alpha_S=0.125$, eqs.~\nr{physMS}
complete the relation of $\overline{\rm MS}$
to Physics.
\subsection{Numerical results for the parameters
of the effective 3d theory in Standard Model}
\label{results}
In Secs.~\ref{ggprim} and~\ref{nulgY} we have expressed
the five parameters $g^2(\mu)$, $g'^2$,
$\nu^2(\mu)$, $\lambda(\mu)$, and $g_Y^2(\mu)$ in terms
of the five physical parameters $G_\mu$, $m_W$, $m_Z$,
$m_H$ and $m_t$. In addition, due to the
large experimental uncertainty in $m_W$,
we replaced $m_W$ as an input
parameter with the hadronic contribution to the photon self-energy
through Table~1.
We have then four parameters left: the very well known
$G_\mu=1.16639\times 10^{-5}$ GeV$^{-2}$~\cite{PDG},
$m_Z=91.1887$~GeV~\cite{C}, the less well
known $m_t=175$ GeV~\cite{CDF}, and the
unknown $m_H$. We shall fix $m_t$ and use the
Higgs mass as a free parameter. Then
we can calculate the parameters of the effective
3d theory in terms of $m_H$ and $T$.
We do not write down the formulas
for the 3d parameters in terms of the physical 4d
parameters from eqs.~\nr{fineq},~\nr{PH}-\nr{PF},~\nr{physMS}
and Secs.~\ref{ioss},~\ref{iohs} explicitly,
since we found no significant simplification in the final result
apart from the $\ln\mu$-terms.
Numerically,
the properties of the 3d SU(2)$\times$U(1)+Higgs theory relevant
for the EW phase transition can be presented
as a function of the physical parameters through a few figures.
First, put $m_Z\to m_W$ so that $g'=0$. Then the final 3d
theory has three parameters: the scale is given
by $\bar{g}_3^2$, and the dynamics is given by the two dimensionless
ratios $x=\bar{\lambda}_3/\bar{g}_3^2$,
$y=\bar{m}_3^2(\bar{g}_3^2)/\bar{g}_3^4$.
The scale $\bar{g}_3^2$ is given as a function
of $m_H$ and $T$ in Fig.~\ref{g3mHT}, and the
parameters $x$ and $y$ are given in Fig.~\ref{xymHT}.
The phase diagram of the theory with the
parameters $x$, $y$,
together with the values of the latent heat,
surface tension and correlation lengths in units of $\bar{g}_3^2$,
have been studied with lattice MC simulations
in~\cite{FKLRS}.
Finally, one has to account for
the effect of the U(1)-subgroup
on the EW phase transition. Since
there are no lattice simulations available for the 3d
SU(2)$\times$U(1)+Higgs model, the best one can do is
to estimate the effect of the U(1)-subgroup
perturbatively. In Fig.~\ref{U1}, we display the
percentual perturbative effect of the U(1)-subgroup on the
critical temperature $T_c$, the vacuum
expectation value of the Higgs field~$v$, the latent heat $L$,
and the surface tension $\sigma$ as a function of the physical Higgs mass.
Using the non-perturbative
values for the case $g'=0$ from~\cite{FKLRS}, one can then
derive results for the full Standard Model.
\subsection{The effect of higher-order operators}
\label{corrections}
In Landau gauge in the Standard Model,
the dominant 6-dimensional $\phi^6$-operators related to
the integration over the superheavy scale are
\begin{eqnarray}
O^{(6)}_{g^2} & = & \frac{3\zeta(3)}{16\,384\pi^4}\frac{g^6\phi^6}{T^2},
\label{o6g2} \\
O^{(6)}_{g_Y^2} & = & -\frac{7\zeta(3)}{512\pi^4}\frac{g_Y^6\phi^6}{T^2}.
\label{o6gY}
\end{eqnarray}
These follow from eq.~\nr{hoJ} in Sec.~\ref{blocks}.
A complete list of the dominant fermionic
contributions to the other 6-dimensional
operators has been worked out in~\cite{M}.
The dominant $\phi^6$-operator related to the integration
over the heavy scale is
\begin{equation}
O^{(6)}_{\rm heavy}=\frac{3\sqrt{6/5}}{10\,240\pi}\frac{g^3\phi^6}{T^2}.
\label{o6h}
\end{equation}
The 6-dimensional operators are neglected in the effective theories
discussed in this paper, and their
importance has to be estimated.
It is rather difficult to estimate
the effect of the 6-dimensional operators
comprehensively, apart from the powercounting estimate
in Sec.~\ref{DR}. What can be done easily, though,
is an evaluation of the shift caused by the $\phi^6$-operators
in the vacuum expectation value of the Higgs field.
A generic 6-dimensional operator $O^{(6)}=c\phi^6/T^2$
produces the term $\delta V=c\varphi^6/T^2$
to the effective potential $V(\varphi)$.
Through
\begin{equation}
V'(\varphi+\delta\varphi)+\delta V'(\varphi)=0,
\end{equation}
the relative shift induced is
\begin{equation}
\frac{\delta\varphi}{\varphi}=-\frac{\delta V'}{\varphi V''}
\approx -\frac{3c}{\lambda}\frac{\varphi^2}{T^2},
\end{equation}
where it was assumed that $V''(\varphi)\sim 2\lambda \varphi^2$.
For $m_H\sim m_W$ so that $\lambda\sim g^2/8$, the coefficients
$3c/\lambda$ in the Standard Model
following from eqs.~\nr{o6g2}, \nr{o6gY},
\nr{o6h} are
\begin{equation}
\biggl(\frac{3c}{\lambda}\biggr)_{g^2}\sim 10^{-5},\quad
\biggl(\frac{3c}{\lambda}\biggr)_{g_Y^2}\sim -10^{-2},\quad
\biggl(\frac{3c}{\lambda}\biggr)_{\rm heavy}\sim 10^{-3}.
\end{equation}
The contributions from the top quark are seen to be
dominant, and in the region where $\varphi/T\sim 1$, the effect
of the corresponding 6-dimensional operator is of the order of one
percent. Note that from the point of view of the 6-dimensional operators,
the integration over the heavy scale is relatively {\em better} than
the integration over the superheavy scale, although in terms
of powers of coupling constants, the accuracy is worse.
We conclude that the final 3d effective theory for the light
fields should give results accurate within a few percent for all
thermodynamic properties of the phase transition, like the
latent heat, the surface tension, and the correlation lengths.
For the critical temperature, the accuracy should be
an order of magnitude better. In the pure SU(2)+Higgs theory
without fermions, the accuracy of the theory with light and heavy
fields should be better than 1\% for all thermodynamic properties.
\section{Discussion}
\label{conclusions}
The set of diagrams described and computed in Section 3 is sufficient
to make a dimensional reduction of a large class of theories.
In particular, it can be used for a construction
of an effective 3d theory for different extensions of the Standard Model.
Below we will argue that in many cases the effective theory appears to
be just the SU(2)$\times$U(1)+Higgs model. We do not attempt to carry out
the necessary computations here and discuss the general strategy only.
Let us take as an example the two Higgs doublet model. The integration
over the superheavy modes gives a 3d SU(2)$\times$U(1)
theory with two Higgs doublets,
one Higgs triplet and one singlet (the last two are
the zero components of the gauge fields).
Construct now the 1-loop scalar mass matrix for the doublets and find
the temperatures
at which its eigenvalues are zero. Take the higher temperature; this is
the temperature near which the phase transition
takes place. Determine the mass of
the other scalar at this temperature. Generally, it is of the order of
$g T$, and therefore, is heavy. Integrate it out together with the
heavy triplet and singlet -- the result is the
simple SU(2)$\times$U(1) model. In the
case when both scalars are light near the critical temperature a
more complicated model, containing two scalar doublets, should be studied.
It is clear, however, that this case requires some fine tuning.
The consideration of the phase transitions in
the two Higgs doublet model on 1-loop
level can be found in~\cite{BKS,TZ}.
The same strategy is applicable to the Minimal Supersymmetric Standard Model.
If there is no breaking of colour and charge at high temperature (breaking
is possible, in principle,
since the theory contains squarks),
then all degrees of freedom, excluding those belonging to the two Higgs
doublet model, can be integrated out. We then return back to the case
considered previously. The conclusion in this case is similar to the
previous one, namely that the phase transition in MSSM can be described
by a 3d SU(2)$\times$U(1) gauge-Higgs model, at least in a part of the
parameter space. A 1-loop analysis of this theory was carried
out in~\cite{Gi,BEQ,EQZ}.
The procedure of dimensional reduction will give an infrared safe connection
between the parameters of the underlying 4d theory and those
of the 3d theory. The latter can then be studied by non-perturbative means,
such as lattice Monte Carlo simulations~\cite{KRS,FKRS,knp,IKP,FKLRS}.
|
\section{\bf Introduction}
Paczynski's bold proposal \cite{pac} to use microlensing to
probe the Galactic halo for dark compact baryonic objects
(referred to as MACHOs for Massive Astrophysical Compact Halo
Objects) has become a reality. Three collaborations, EROS, OGLE
and MACHO have reported over eighty microlensing events towards
the Galactic Bulge and eight in the direction of the Large
Magellanic Cloud (LMC) \cite{eros,ogle,MACHO,MACHObulge}. (Preliminary
analyses of the second year MACHO data toward the LMC indicate
two new events \cite{lmcrumor}.) This detection of microlensing
has opened up a new window for exploring the dark halo of our
galaxy. In this paper we use the existing data to shed light on
the composition of the Galactic halo. Some of our key
results have been summarized elsewhere \cite{prl,apjlett}; here
we present the full details of our analysis.
There is compelling evidence that spiral galaxies are imbedded
in extended non-luminous halos. This includes flat rotation
curves measured for almost 1000 spiral galaxies, studies of
binary galaxies including our own galaxy and M31, weak
gravitational lensing, flaring of neutral hydrogen in the disks
and studies of disk warping \cite{haloevid,bbs}. While the halo of
our own galaxy is in many respects more difficult to study,
there is much important data here too; e.g., the rotation curve
has been measured between 4$\,{\rm kpc}$ and 18$\,{\rm kpc}$, the flaring of
hydrogen gas has been studied, and the orbital motions of
globular clusters and satellite galaxies have been determined
\cite{ourhalo}. All of these support the hypothesis of an
extended dark halo.
Although there is strong evidence for the existence of a
Galactic halo, there is little direct information concerning its
composition. Since the halos of spiral galaxies are large and
show little sign of having undergone dissipation they can be
expected to reflect the composition of the Universe as a whole,
though perhaps with some biasing (severe in the case of hot dark
matter), and thus their composition is of more
universal importance. X-ray observations rule out a hot,
gaseous halo, and the Hubble Space Telescope has placed tight
limits on the contribution of faint stars \cite{hst}. The most
promising candidates for the halo material are baryons in the
form of MACHOs and cold dark matter (CDM) particles.
A baryonic halo invokes the fewest hypotheses: Brown dwarves are
known to exist. Further, substantial baryonic dark matter must
exist given the robust nucleosynthesis lower bound on
$\Omega_B$\cite{turnerschramm}. However, the success of CDM
models in explaining the formation of large-scale structure and
the appeal of a flat universe and the nucleosynthesis bound to
$\Omega_B$ make a strong case for CDM. If the bulk of the
Universe exists in the form of CDM, it is inevitable that our
halo contains a significant CDM component \cite{gates}. (Even
in the most radical scenario for the formation of the Galaxy,
infall onto a baryonic seed mass, the amount of CDM accreted is
at least equal to the total baryonic mass of the galaxy.)
Conclusively demonstrating that the halo is not composed solely
of baryons would comprise additional strong support, albeit
circumstantial, for a halo comprised of CDM particles.
Gravitational microlensing provides a valuable tool for probing
the baryonic contribution to the halo---and of the structure of
the Galaxy itself. We shall focus on measurements of the
optical depth for microlensing (the probability that a given
distant star is being microlensed). The optical depth is
determined by the amount and distribution of mass in microlenses
along the line of sight. With sufficient lines of sight a sort
of galactic tomography could in principle be performed. At
present only a few lines of sight have been probed: several in
the direction of the LMC, which probe the halo, and several in
the direction of the Bulge, which probe the inner galaxy. The
small probability for microlensing, of order $10^{-6}$, means
that millions of stars must be monitored. There are many fields
of view available in the direction of the Bulge and so
tomography of the inner galaxy is a realistic possibility. The
situation for probing the halo is not as promising: with available resources
only the direction toward the LMC has star
fields of sufficiently high density to be useful. However, a
space-based search should be able to target the Small Magellanic
Cloud (SMC), perhaps some of the larger globular clusters and
the closer galaxies such as M31.
Even with precise knowledge of the optical depths toward the LMC
and bulge, it would still be difficult to interpret the results
because of the large uncertainties in the structure of the
Galaxy. As it is, small number statistics for the LMC lead to a
range of optical depths further complicating the analysis.
Detailed modeling of the Galaxy is essential to drawing reliable
conclusions.
Thus, we adopt the following strategy for determining the MACHO
composition of our galactic halo. We construct models of the
Galaxy with five components: luminous and dark disks, baryonic
and CDM halos, and a bulge. We describe each by parameters
whose values are allowed to vary over a range motivated by
previous modeling and observations. By simultaneously varying
all the parameters we construct a very large space of models
(more than ten million); from this we find a subspace of viable
models consistent with the diverse set of observations that
constrain the Galaxy---rotation curve, local projected mass
density, measurements of the the amount of luminous matter in
the disk and bulge, and measurements of the optical depth for
microlensing toward the bulge and the LMC. The distribution of
the MACHO halo fraction in these viable models allows us to
infer its preferred value. Further, since it is difficult to
exclude an all-MACHO halo we focus attention on models where the
MACHO fraction is high to see what observations might be crucial
in testing this possibility.
Our approach is not the only one that could be pursued. The
MACHO Collaboration has focused on a handful of representative
galactic models that are meant to span the larger range of
possibilities \cite{newMACHO}. This allows them to study each
model in more detail and address not only the number of
microlensing events, but also their durations (which are
determined by a combination of the MACHO mass, distance and
velocity across the sky). They reach a similar conclusion
concerning the MACHO fraction of the halo---it is small in most
models of the Galaxy---though they construct a model with an
all-MACHO halo. While their approach allows them to address the
question of the masses of MACHOs, they do not constrain their
models with the totality of observations and thus they cannot
address the viability of the models they consider. Indeed, we
find their all-MACHO model incompatible with the observational
data.
A few caveats should be kept in mind. Because the acceptance of
the MACHO and EROS experiments to event duration are limited,
the present data address only the halo component made up of
MACHOS with masses from about $10^{-7}M_\odot$ to $10^2
M_\odot$. It has been argued that objects of mass outside this
range are unlikely: MACHOs of mass $10^{-7} M_\odot$ evaporate
on a time scale less than the age of the galaxy \cite{derujula};
Black holes of greater than $10^4 {M_\odot}$ would disrupt the
globular clusters \cite{carr}. However, there remains the
possibility that the halo baryons are in the form of either
molecular clouds with a fractal distribution \cite{depaolis} or
very massive ($m \sim 10^2 M_\odot - 10^4 M_\odot$) black holes
\cite{carr}. Neither of these options is particularly
compelling---molecular clouds should have collapsed by the
present and the massive progenitors of such black holes would
likely have produced $^4$He or heavy elements---however, they
cannot be ruled out conclusively at this time.
In our analysis we also assume that MACHOs are smoothly
distributed rather than clumped. If they were strongly clumped
the microlensing rate could vary significantly across the sky,
which might appear to allow a smaller or larger optical depth
toward the LMC for a given MACHO halo fraction. However, if
more than one clump were on average expected in a patch of sky
the size of the LMC then the optical depth would be again close
to its average. Thus, for clumping to significantly affect the
optical depth there must be at most a few clumps in the solid
angle subtended by the LMC. But if this is the case, then we
can expect no more than a few thousand such clumps over the
entire sky out to the distance of the LMC. To be a significant
fraction of the total halo mass ($\sim {\rm few} \times 10^{11}
M_{\odot}$) each clump must be of order ${\rm few} \times 10^{8}
M_{\odot}$, far greater than the mass of a globular cluster. A
few thousand of these objects residing in the halo would seem to
be ruled out firmly by dynamical constraints based upon the
stability of the disk \cite{carr}.
Our paper is organized as follows: In the next Section we
discuss galactic modeling and the minimal constraints we impose
on models. In Section 3 we discuss the implications of
microlensing on galactic modeling. We also consider additional
reasonable constraints, the local escape velocity and satellite
galaxy proper motions, which preclude any model with an
all-MACHO halo. In Section 4 we examine more closely the few
models that allow an all-MACHO halo (within the minimal
constraints) as well as those models that allow a no-MACHO halo.
In the final Section we summarize our results and discuss future
observations---from measurements of galactic parameters to
strategies for the microlensing measurements---that can sharpen
conclusions concerning the MACHO fraction of the halo.
\section{\bf Galactic Modeling}
Modeling of the Galaxy is an established subject---the basic
features and dimensions of the Galaxy were determined early in
this century---but also one that is still undergoing significant
change. Evidence for a dark halo has accumulated over the past
two decades (see, e.g. \cite{glxmodels}) and over the past five
years or so a strong case has for a bar-like, rather than
axisymmetric, bulge has developed \cite{manybars}. Microlensing
has the potential for contributing significantly to our
understanding of the structure of the Galaxy, both of the
composition of the halo and the mass distribution interior to
the solar circle.
The current picture of the Galaxy is a barred spiral,
consisting of three major components: a central bulge (bar), a
disk and a dark halo. The luminous components are a thin,
double exponential disk with a vertical scale height of about
$0.3\,{\rm kpc}$ and a radial scale length of about $3.5\,{\rm kpc}$, a
smaller (few percent of the disk mass) ``thick'' disk with
vertical scale height of about $1\,{\rm kpc}$ to $1.5\,{\rm kpc}$ \cite{GWK},
and a central bulge region, which recent observations indicate
is a triaxial bar \cite{manybars}.
Evidence for the dark halo is less direct, but firm nonetheless.
It comes from the rotation curve, which is flat out to at least
$18\,{\rm kpc}$ (and probably out to $50\,{\rm kpc}$) and the approach of
Andromeda and the Galaxy toward one another. At the solar
circle about 40\% of the centripetal acceleration is provided by
the gravitational force of the halo, and beyond that the
fraction is even greater. The mass of the Galaxy inferred from
the approach of Andromeda is at least a factor of ten greater
than that which can be accounted for by stars alone
\cite{haloevid}. Moreover, the evidence for dark halos
associated with spiral galaxies in general is very secure. A
recent survey of the rotation curves of more than 900 spiral
galaxies indicates flat or slightly rising rotation curves at
the limit of the observations, providing strong evidence for
their massive dark halos \cite{persic2}. {}From a completely
different direction, Brainerd, Blandford and Smail \cite{bbs}
have mapped the dark halos of several spiral galaxies by means
of their weak-gravitational lensing of very distant galaxies.
Their results indicate that the halos studied have radial extent
of at least $100 h^{-1}\,{\rm kpc}$ and total masses in excess of
$10^{12} M_{\odot}$.
The values of the parameters that describe the components of the
Galaxy are not well determined; this is especially true for the
halo whose presence is only known by its gravitational effects.
In addition, there is interplay between the various components
as the observations typically constrain the totality of the
model, rather than a given component. Modeling uncertainties
introduce significant, irreducible uncertainties in the
determination of the MACHO content of the halo. In order to
understand these uncertainties we explore a very wide range of
models that are consistent with all the data that constrain the
Galaxy.
We consider two basic models for the bulge, the first following
Dwek et al. \cite{dwek} who have utilized DIRBE surface
brightness observations to construct a triaxial model for the
bulge:
\begin{equation} \rho_{\rm BAR} = {M_0 \over 8 \pi abc}
e^{-s^2/2}, \qquad s^4 = \left[ {x^2\over a^2} + {y^2\over b^2}
\right] ^2 + {z^4 \over c^4} ,
\end{equation}
where the bulge mass $M_{\rm Bulge} = 0.82 M_0$, the scale
lengths $a = 1.49\,{\rm kpc}$, $b = 0.58\,{\rm kpc}$ and $c = 0.40\,{\rm kpc}$, and
the long axis is oriented at an angle of about $10^{\circ}$ with
respect to the line of sight toward the galactic center. While
we do not take the axes and inclination angles to be modeling
parameters, we later explore the sensitivity of our results to
them. We also consider an axisymmetric Kent model for the bulge
\cite{kent}. The rotation curve contribution was calculated in
the point mass approximation. At $r=5\,{\rm kpc}$ this approximation is accurate to
better than $10\%$.
The bulge mass is not well determined, and we consider $M_{\rm
Bulge}=(1 - 4) \times 10^{10} M_\odot$, in steps of $0.5\times
10^{10}M_\odot$. Previous estimates have been in the range $(1
- 2) \times 10^{10} M_\odot$ \cite{glxmodels,kent,zhao},
although a recent study by Blum \cite {blum} which utilized the
tensor virial theorem found a bar mass closer to $3 \times
10^{10} M_\odot$ (assuming a bar orientation of 20 degrees --
smaller (larger) angles of orientation imply larger (smaller)
bulge masses).
For the disk component we take the sum of a ``fixed,'' thin
luminous disk whose constituents (bright stars, gas, dust, etc.)
are not expected to serve as lenses,
\begin{equation}
\rho_{\rm LUM}(r,z) = {\Sigma_{\rm LUM}\over 2h}\,
\exp [-(r-r_0)/r_d] e^{-|z|/h},
\end{equation}
with scale length $r_d = 3.5\,{\rm kpc}$, scale height $h = 0.3\,{\rm kpc}$,
and local projected mass density $\Sigma_{\rm LUM} = 25
M_\odot\,{\rm pc}^{-2}$ \cite{lumdisk}, and a ``variable'' disk
component whose constituents are assumed to be lenses. For the
variable component we consider first a distribution similar to
that of the luminous matter but with varying scale lengths $r_d
= 3.5\pm 1 \,{\rm kpc}$, and thicknesses $h=0.3\,{\rm kpc}$, and $1.5\,{\rm kpc}$.
We also consider a model where the projected mass density varies
as the inverse of galactocentric distance (Mestel model)
\cite{1/r}.
The motions of stars perpendicular to the galactic plane have been
used to infer the {\it total} local projected mass density
within a distance of $0.3\,{\rm kpc} - 1.1\,{\rm kpc}$ of the galactic plane
\cite{sigma0}. The values so determined are between
$40M_\odot\,{\rm pc}^{-2}$ and $85M_\odot\,{\rm pc}^{-2}$. As a
reasonable range we require that $\Sigma_{\rm TOT}(1\,{\rm kpc} ) =
\int_{-1\,{\rm kpc}}^{1\,{\rm kpc}}\rho (r_0, z)dz = 35 - 100 M_\odot\,{\rm
pc}^{-2}$, which constrains the local projected mass density
of the dark disk to be $10M_\odot\,\le \Sigma_{\rm VAR}
\le 75 M_\odot \,{\rm pc}^{-2}$. (We also include
the contribution of the halo to $\Sigma_{\rm TOT}(1\,{\rm kpc} )$,
which for flattened halo models can be significant, about $20
M_\odot\,{\rm pc}^{-2}$, and reduces the mass density that the
variable disk can contribute.)
The dark halo is assumed to be comprised of two components,
baryonic and non-baryonic, whose distributions are independent.
We first assume independent isothermal distributions for MACHOs
and cold dark matter with core radii $a_i = $2, 4, 6, ..., 18,
20$\,{\rm kpc}$,
\begin{equation}
\rho_{{\rm HALO},i} = {a_i^2+r_0^2 \over a_i^2 +r^2}\, \rho_{0,i}\ ,
\label{eq:iso}
\end{equation}
where $i=$ MACHO, CDM and $\rho_{0,i}$ is the local
mass density of component $i$.
There are indications from both observations \cite{Sackflat,rix}
and CDM simulations \cite{quinn} that halos are significantly
flattened. In order to explore the effects of flattening we
also consider models with an axis ratio $q = 0.4$ (an E6 halo)
for both the baryonic and non-baryonic halos with distributions
of the form
\begin{equation}
\rho_{{\rm HALO},i} = {a_i^2+R_0^2 \over a_i^2 +R^2 + (z/q)^2}\,
\rho_{0,i}\ ,
\end{equation}
where $(R,z)$ are cylindrical coordinates. While flattening
does affect the local halo density significantly, increasing it
by roughly a factor of $1/q$ (see Ref.~\cite{apjlett}), it does not
affect the halo MACHO fraction significantly.
Finally, we consider the possibility that the MACHOs are not
actually in the halo, but instead, due to dissipation, are more
centrally concentrated. To describe this we use the
distribution in Eq.~(\ref{eq:iso}) but with $r^2$ replaced by
$r^n$, for $n=3,4$ and core radii $a_{\rm MACHO} =1,2 \,{\rm kpc}$.
Such a distribution approximates models of a spheroidal
component \cite{glxmodels,giudice} (note, in these models we
also explicitly include a Dwek bar).
We construct our models of the Galaxy by letting the parameters
describing the various components vary independently. By doing
so we consider millions of models. We pare down the space of
models to a smaller subset of viable models by requiring that
observational constraints be satisfied. The kinematic
requirements for our viable models are: circular rotation speed at
the solar circle ($r_0 = 8.0\,{\rm kpc} \pm 1\,{\rm kpc}$) $v_c =220\,{\rm km\,s^{-1}}\pm
20\,{\rm km\,s^{-1}}$; peak-to-trough variation in $v(r)$ between $4\,{\rm kpc}$ and
$18\,{\rm kpc}$ of less than 14\% (flatness constraint \cite{gates});
and circular rotation velocity at $50\,{\rm kpc}$ greater than
$150\,{\rm km\,s^{-1}}$ and less than $307\,{\rm km\,s^{-1}}$. We first impose this minimal set
of constraints in order to be as conservative as possible
in our conclusions; later we impose additional reasonable, but
less secure constraints, involving the rotation curve at large
distances and the local escape velocity.
We also impose constraints from microlensing, both toward the
bulge and toward the LMC. The optical depth for microlensing a
distant star by a foreground star is \cite{griest}
\begin{equation}
\tau = {4 \pi G\over c^2} {\int^\infty_0 ds \rho_s (s)
\int^s_0 dx \rho_l (x) {x(s-x)/ s} \over \int^\infty_0 ds \rho_s (s)},
\end{equation}
where $\rho_s$ is the mass density in source stars, $\rho_l$ is
the mass density in lenses, $s$ is the distance to the star
being lensed, and $x$ is the distance to the lens \cite{lenses}.
In calculating the optical depth toward the bulge, we consider
lensing of bulge stars by disk, bulge and halo objects; for the
LMC we consider lensing of LMC stars by halo and disk objects.
Except where we are constructing microlensing maps of the bulge
(see Section 5) we define the direction of the bulge to be toward
Baade's window, $(b,l) = (-4{^\circ},1{^\circ})$.
We adopt the following constraints based upon microlensing data: (a)
$\tau_{\rm BULGE} \geq 2.0\times 10^{-6}$ and (b) $0.2\times
10^{-7}\le \tau_{\rm LMC} \le 2\times 10^{-7}$ \cite{larger}.
The bulge constraint is based upon the results of the OGLE
Collaboration \cite{ogle} who find $\tau_{\rm BULGE} = (3.3\pm
1.2)\times 10^{-6}$, as well as the results of the MACHO
Collaboration who find $\tau_{\rm BULGE} = 3.9 ^{+1.8}_{-1.2}\times 10^{-6}$
\cite{MACHObulge}. To be sure, there are still important
uncertainties, e.g., detection efficiencies and whether or not
the stars being lensed are actually in the bulge; however, we
believe this to be a reasonable bound to the optical depth. The
optical depth to the LMC is based upon the MACHO Collaboration's
measurement \cite{MACHOprl}, $\tau_{\rm LMC} = 0.80\times 10^{-7}$,
as well as the results of the EROS Collaboration \cite{eros}.
Here too there are uncertainties. In addition to the obvious
small number statistics, the events might not all be
microlensing. As a reasonable first cut we have taken the 95\%
Poisson confidence interval based upon the MACHO results.
Bulge microlensing provides a crucial constraint to galactic
modeling and eliminates many models. It all but necessitates a
bar of mass at least $2\times 10^{10}M_\odot$, and, as has been
emphasized by others \cite{zhao}, provides additional evidence
that the bulge is bar-like. Because of the interplay between
the different components of the Galaxy, the bulge microlensing
optical depth indirectly constrains the MACHO fraction of the halo.
On the other hand, LMC microlensing only constrains the MACHO
fraction of the halo.
\section{\bf Implications of Microlensing for Galactic Modeling}
In this Section we discuss the characteristics of the viable
models, focussing particularly on the composition of the halo
(MACHO fraction and local halo mass density), but also paying
attention to the other parameters in our galactic models. We
display our results in histograms of the number of viable models
as a function of various modeling and derived parameters.
These plots {\it resemble} likelihood functions that are
marginalized with respect to those parameters. They are in fact
not likelihood distributions; because the most important
uncertainties in modeling the Galaxy are systematic in
character, e.g., the model of the Galaxy itself, the rotation
curve, the shape of the halo, and even the galactocentric
distance and local speed of rotation, we resisted the urge to
carry out a more rigorous statistical analysis which might have
conveyed a false level of statistical significance.
We first discuss the features of the models that satisfy our
minimal constraints and then go on to discuss the models that
survive when we impose additional constraints that better serve
to define the extent of the dark halo (escape velocity and
rotation curve at large distances as defined by satellite galaxy
proper motions). In these discussions we rely heavily upon
histograms which detail the characteristics of the acceptable
galactic models. However, before we do, let us summarize our
main results:
\begin{itemize}
\item In most viable models the halo MACHO fraction is between
0\% and 30\%, though when only the minimal constraints are
applied there are models with MACHO fraction greater than
60\%. When the additional constraints
are applied there are no viable models with
halo MACHO fraction greater than 60\% (see Fig.~1).
(Halo MACHO fraction $f_B$ is defined to be the MACHO mass
fraction of the halo interior to $50\,{\rm kpc}$).
\item In viable models the local MACHO mass density is
sharply peaked around $10^{-25}{\,{\rm g\,cm^{-3}}}$ (see Fig.~2) and the total
MACHO mass (within $50\,{\rm kpc}$) is peaked around $1\times 10^{11}
M_\odot$.
\item In viable models with a flattened halo
the total local halo mass density
is between about $4\times 10^{-25}{\,{\rm g\,cm^{-3}}}$ and
$1.5\times 10^{-24}{\,{\rm g\,cm^{-3}}}$ (see Fig.~2). Flattening
increases the local halo mass density by factor of
order the axis ratio.
\item The bulge microlensing constraint precludes any model with
a Kent (axisymmetric) bulge, and the bar mass in most viable
models is between $2\times 10^{10}M_\odot$ and $3\times
10^{10}M_\odot$. The necessity of a relatively heavy galactic
bar plays an important role constraining the halo MACHO fraction
to a small fraction.
\end{itemize}
\subsection{Minimal constraints}
There are several features that are generic to most models that
satisfy the minimal set of constraints (see Figs.~3-8). The
most important of these is that independent of almost all the
model parameters, the peak of the MACHO fraction occurs for $f_B
\mathrel{\mathpalette\fun <} 20\%$ (the only exception being a spherical halo model with
a very small core radius for the non-baryonic component, which
peaks at $f_B \sim 30\%$). While the range of MACHO fraction
extends from 0\% to 90\%, most models have $f_B < 30\%$. (We
discuss the handful of high MACHO-fraction models in the next
Section). No model with a thick dark disk (either exponential
or $1/r$ profile) and $f_B > 60\%$ survives our constraints, and
the distribution for these thick disk models peaks at $f_B \sim
0$. The absence of MACHOs in the halo is allowed because a
thick disk can contribute up to $0.5\times 10^{-7}$ to the
optical depth toward the LMC \cite{prl}, which allows the LMC
microlensing constraint to be satisfied without recourse to
MACHOs in the halo.
The bulge mass in most models is between $2\times 10^{10} M_\odot$
and $3\times 10^{10}M_\odot$, which is consistent with estimates
from recent efforts to model the bar \cite{zhao,blum}. Models with a Kent
bulge do not provide sufficient microlensing toward the bulge, and as pointed
out in previous work by the authors and others \cite{prl,goulddisk}, the disk
cannot provide more than about $1 \times 10^{-6}$ to the optical depth toward
the bulge. A heavy
bar is necessary to obtain optical depths to the galactic bulge
in excess of $3 \times 10^{-6}$, as currently suggested by the
experimental data. The distribution of galactocentric distance
($r_0$) is somewhat dependent on the disk model, with thick disk
models generally favoring smaller $r_0$. The distribution for
the local circular velocity is relatively broad, but it is
generally peaked at the low end of the range, around $210\,{\rm km\,s^{-1}}
-220 \,{\rm km\,s^{-1}}$. The trend for all dark disk models is toward larger
scale length (r$_d$). The value of the disk surface density
depends on the disk model, although lighter disks are favored in
all cases (i.e., little mass in the dark disk).
The distribution of optical depths toward the LMC and the bulge
are shown in Figs.~3-8. In general, $\tau_{\rm LMC}$ is
relatively flat. This is easily understood: for a given model,
the microlensing optical depth is sensitive only to the MACHO
fraction, which is unaffected by the kinematic cuts. For
thick-disk models (both exponential and $1/r$) there is also a
relatively large bin at the smallest allowed value of $\tau_{\rm
LMC}$. This is due to additional allowed models with very small
halo MACHO fraction where the LMC lensing is done by the disk
(lensing toward the LMC is negligible in thin-disk models
\cite{prl}). The bulge optical depth is somewhat peaked toward
the low end of the acceptable range, mainly due to the
difficulty of achieving $\tau_{\rm bulge} > 3 \times 10^{-6}$.
The local MACHO mass density peaks at about $10^{-25}
{\,{\rm g\,cm^{-3}}}$ in all models and the mass of MACHOs in the
halo peaks at about $1\times 10^{11}M_\odot$. However, the total local halo
mass density is more dependent on the halo model, in particular on
whether or not the halo is flattened; see Fig.~2. (Since the MACHO fraction
of the halo is small, this also applies to the local mass
density of CDM particles.) Flattening of the halo, for
which there is good evidence, increases the local halo
density by a factor of order the axis ratio $q$. In a flattened halo model,
the local halo density is larger by a factor
\begin{equation}
{\rho_0^{flattened} \over \rho_0^{spherical}} = {\sqrt{1-q^2}\over q \sin^{-1}
(\sqrt{1-q^2}) },
\end{equation}
relative to a spherical halo model with the same asymptotic rotation velocity
and core radius (for the E6 halo,
this factor is about 2.)
This has important implications for the
direct detection of non-baryonic dark matter, and is discussed
in detail elsewhere \cite{apjlett}. However, our results for the MACHO
fraction of the halo are essentially independent of the amount of halo
flattening as can be seen in figures 3-11. Both the total mass of the halo and
the MACHO halo mass shift slightly toward smaller values in a flattened halo
model.
\subsection{Additional constraints}
The models we have considered viable thus far have been subject
to a very minimal set of constraints -- that is, we have
tried to be as generous as possible in admitting models, probably
too generous. There are additional constraints which bear on the
size and extent of the dark halo. They are especially crucial
to the issue of the MACHO fraction of the halo: Microlensing
toward the LMC closely constrains the mass of the MACHOs in the
halo, and therefore the halo MACHO fraction depends sensitively
upon the total halo mass. The models with high MACHO fraction
are characterized by light halos; the additional constraints
place a stringent lower bound to the halo mass and thus upper
bound to the MACHO fraction, eliminating all models with MACHO
fraction greater than 60\%.
The first additional constraint on the galactic potential that
we consider comes from the local escape velocity. Based upon
the velocity of the fast moving stars Leonard and Tremaine
\cite{leonard} have determined that the local escape velocity
lies in the range $450 \,{\rm km\,s^{-1}} < v_{\rm ESC} < 650\,{\rm km\,s^{-1}}$ (with 90\%
confidence level), with a stronger lower limit of $430\,{\rm km\,s^{-1}}$.
Kochanek \cite{kochanek} obtains a slightly higher range of $489
\,{\rm km\,s^{-1}} < v_{\rm ESC} < 730\,{\rm km\,s^{-1}}$. Based on these values we adopt
$v_{\rm ESC} > 450\,{\rm km\,s^{-1}}$.\footnote{The escape velocity from an
isothermal halo increases logarithmically; to compute $v_{\rm
ESC}$ we truncate the halo at a distance of $100\,{\rm kpc}$.}
Next we consider the information about the galactic rotation
curve at large distances ($50\,{\rm kpc} - 100\,{\rm kpc}$) based upon the
proper motions of satellites of the Milky Way. Recently Jones,
Klemola and Lin \cite{Jones} have measured the proper motion of
the LMC. They find a total galactocentric transverse velocity
of $215 \pm 48\,{\rm km\,s^{-1}}$. Proper motions for Pal 3 \cite{Cudworth}
(galactocentric distance $79\,{\rm kpc}$) and Sculptor
\cite{Schweitzer} (galactocentric distance $95\,{\rm kpc}$) have also
been measured, yielding $252\,{\rm km\,s^{-1}} \pm 85\,{\rm km\,s^{-1}}$ and $199\,{\rm km\,s^{-1}} \pm
58\,{\rm km\,s^{-1}}$ respectively. Assuming that these satellite galaxies
are bound to our Galaxy, they provide strong evidence that the
galactic halo is massive and extended.
Finally, a study of the rotation curves of over 900 spiral
galaxies \cite{persic2} indicates that for all of these galaxies
the rotation curves are flat, rising or only gently falling at
twice the optical radius ($r_{\rm opt}\equiv 3.2r_d$), depending
on the luminosity. Based on rotation curves of galaxies similar
to the Galaxy ($L/L_*=1.4h^{2}, r_d \approx 3.5\,{\rm kpc}$), the
rotation velocity at $2r_{\rm opt} \sim 22\,{\rm kpc}$ should be within
a few percent of $v_c$, and further, at a galactocentric
distance of $50\,{\rm kpc}$ the rotation velocity should be at least
$200\,{\rm km\,s^{-1}}$. Combining this with the satellite proper motions we
require $180\,{\rm km\,s^{-1}} \leq v_c(50\,{\rm kpc}) \leq 280\,{\rm km\,s^{-1}}$.
We impose these additional constraints on our ``canonical''
model---E6 halo, thin, double-exponential disk, and Dwek
bar---with all other parameters allowed to vary as before. The
results are displayed in Fig.~9. The most striking consequence
of the additional kinematic constraints is the exclusion of all
models with a baryon fraction greater than 60\%, and essentially
all models with a baryon fraction greater than 50\%. It is
worth noting that this result follows from either constraint
alone. That is, models with an all-MACHO halo are characterized
by {\em both} $v_{\rm ESC}< 450\,{\rm km\,s^{-1}}$ {\em and} $v_c(50\,{\rm kpc})<
180\,{\rm km\,s^{-1}}$. The results for a spherical halo are similar.
The halo MACHO fraction for these models is strongly peaked
around 10\% to 20\%. This result is independent of the bar mass,
local disk surface mass density, disk scale length and our
galactocentric distance. It is also insensitive to the optical
depth for microlensing toward the galactic bulge. It is, as one
would expect, sensitive to the optical depth for microlensing
toward the LMC.
These additional constraints also narrow the estimate for the
total mass of the Galaxy (within $50\,{\rm kpc}$) to $(5 \pm 1) \times
10^{11} M_{\odot}$. This is consistent with the value obtained
recently by Kochanek \cite{kochanek}, who used similar
constraints on the extent of the dark halo, although a much more
restricted set of galactic models.
\section {\bf Very MACHO and No-MACHO Halos}
\subsection{Very-MACHO halos}
In Figs.~3 to 8 the characteristics of galactic models with
MACHO fraction $f_B \ge 0.75$ are shown as dotted lines. (It
should be noted that the histograms for these models with
very-MACHO halos have been multiplied by a factor of 50 relative
to the other models.) The crucial common feature of very-MACHO
models is a light halo (total mass less than $4\times
10^{11}M_\odot$). Only thin-disk models allow $f_B\ge 0.75$.
The reason for this illustrates how the bulge microlensing
constraint also influences other aspects of the galactic model.
Models with an exponential thick disk require a heavier bar to
account for microlensing toward the bulge: A thick disk
contributes far less to microlensing toward the bulge than does
a thin disk \cite{prl}. On the other hand, the rotation curve
from our position outward requires a heavy disk for support if
the halo is light. Therein lies the rub: the inner part of the
rotation curve cannot tolerate both a heavy disk and a heavy
bar.
Because very-MACHO models are characterized by light halos they
are also characterized by: (i) a small local rotation speed,
$v_c \le 215\,{\rm km\,s^{-1}}$; (ii) large (total) local surface mass
density, $\Sigma_{0} \geq 60 M_\odot\,{\rm pc}^{-2}$; (iii)
light bar, $M_{Bulge} =2.0 \times 10^{10} M_\odot$ in most of
these models; (iv) a rotation curve that falls to a small
asymptotic value, $v(50 \,{\rm kpc}) \mathrel{\mathpalette\fun <} 180\,{\rm km\,s^{-1}}$; and (v) a local
escape velocity that is less than $420\,{\rm km\,s^{-1}}$. Further, because
the bar is the most efficient source of lensing, a lighter bar
results in a low optical depth toward the bulge, $\tau_{\rm
bulge}\simeq 2 \times 10^{-6}$. Finally, to avoid having a halo
that is too light, these models are necessarily characterized by
high optical depth toward the LMC, $\tau_{\rm LMC}\sim 2 \times
10^{-7}$.
\subsection{No-MACHO halos}
Because the optical depth for microlensing toward the LMC is so
much smaller than it would be for an all-MACHO halo one should
also consider the possibility that there are no MACHOs in the
halo. Further, the optical depth toward the LMC is based on
only three events seen by the MACHO Collaboration and two by the
EROS Collaboration. Not only are the numbers small, so that
Poisson fluctuations alone are large, but it is not impossible
that some of the events are not even due to microlensing. In
that regard, the MACHO Collaboration refers to their events as
two candidates and one microlensing event (the amplitude 7
event) \cite{newMACHO}, while the EROS Collaboration has established
that one of their events involves a binary star (of period much
shorter than the event duration) \cite{erosbin}. Thus, the actual
optical depth could be quite small.
If the optical depth for microlensing toward the LMC is much
less than $10^{-7}$ (the current central value), it could be
explained by a combination of microlensing of LMC stars by LMC
stars \cite{lmclens} and a thick disk component (a thick disk
can contribute up to $0.5\times 10^{-7}$, though it should
be noted that a thick disk cannot also account for the large
microlensing rate toward the bulge).
Another possibility is that the MACHOs responsible for microlensing
toward the LMC are in a more centrally condensed component, e.g.,
the spheroid. In Figs.~10 and 11 we show the characteristics
of models with a no-MACHO halo and MACHO spheroid with density
profiles $r^{-n}$ ($n=3,4$) and core radii $b=1,2\,{\rm kpc}$. The
viable models are characterized by: (i) very small MACHO fraction,
spheroid mass/halo mass less than 0.2; (ii) very low optical
depth, $\tau_{\rm LMC} \mathrel{\mathpalette\fun <} 5\times 10^{-8}$; and (iii) spheroid
mass which peaks at $5\times 10^{10}M_\odot$ for $n=3$ and $3\times
10^{10}M_\odot$ for $n=4$, consistent with
independent dynamical measurements \cite{giudice}.
\section{Discussion and Summary}
\subsection{\bf Microlensing and the bulge}
The number of microlensing events detected in the direction of
the galactic bulge is currently more than eighty and will continue to
grow. As the statistics improve, the optical depth along different
lines of sight toward the bulge can be determined,
allowing tomography of the inner galaxy, in turn providing
information about the shape, orientation, and mass of the bulge,
and indirectly about the Galaxy as a whole.
Already the unexpectedly high optical depth towards the
galactic center provides further evidence that the bulge is more
bar-like than axisymmetric. Much more can be learned. In
Fig.~12 we present microlensing maps of the bulge for several
different models. The first panel shows contours of constant
$\tau_{\rm bulge}$ for a massive ($M_{Bulge} = 4.0\times
10^{10}M_\odot$) Kent bulge with a light disk. Even with this
very high bulge mass, the microlensing rates are not high enough
to account for the observations. The second panel shows a
microlensing map for a slightly less massive ($M_{Bulge} = 3.0\times
10^{10}M_\odot$) Dwek bar oriented almost directly towards us,
$\theta = 10{^\circ}$. Despite the lower mass which makes this
model more likely to pass kinematic cuts, the optical depths are
much higher, with bulge-bulge events clearly dominating. A
slight asymmetry in galactic longitude is apparent, but it may
be too small to be detected. The third panel shows a microlensing
map for the same mass bar, but oriented at $45{^\circ}$. The
optical depths for microlensing are much smaller, the contours
are considerably less steep and more elongated along the
longitude axis. For comparison, the the effect of a heavier
disk is shown in panels four and five, for a models similar to
those in panels two and three. The additional microlensing
provided by the disk results in higher optical depths and an
elongation of the microlensing contours along the direction of
galactic longitude.
While the orientation of the bar provides a strong signature in
the microlensing maps, the overall rate is an important
constraint by itself. The models shown in panels three and five
with an orientation of $45^\circ$ are already excluded by our
constraint, $\tau_{\rm bulge}\geq 2.0\times 10^{-6}$. Figure 13a
shows the number of viable models with a thin disk and flattened
halo as a function of bar orientation. Clearly a bar pointing
towards us is preferred, with bar orientations of greater than
$30{^\circ}$ almost entirely excluded.
The modeling we have described here has already indicated the
necessity of a relatively massive bar, $(2-3)\times 10^{10}
M_\odot$, even in the case of a bar oriented at $10{^\circ}$ from
our line of sight. This, together with the results shown in
Fig.~13 suggest that the bar has a mass of $(2-3)\times
10^{10}M_\odot$ and is oriented at an angle of less than $20 -
30{^\circ}$ from our line of sight. As discussed earlier,
considering rate alone there is a degeneracy between bar mass
and orientation: lower mass can be traded for smaller angle. As
can be seen in Fig.~12 mapping can break this degeneracy.
\subsection{\bf Future directions}
While the results of the microlensing experiments to date seem
to strongly indicate that the primary component of the halo is
not MACHOs, as we have emphasized here it is not yet possible to
exclude this hypothesis with any certainty. Since the question
is of such importance, it is worth considering future measurements
that could led to more definite conclusions. Based upon our
extensive modeling we can identify a a number of key measurements.
Recall that the models with all-MACHO halos had a number of
distinctive features: (i) large optical depth toward the LMC,
$\tau_{\rm LMC} \simeq 2\times 10^{-7}$; (ii) small optical
depth toward the bulge, $\tau_{\rm bulge}\simeq 2 \times
10^{-6}$; (iii) a small local rotation speed, $v_c \le 215\,{\rm km\,s^{-1}}$;
(iv) large (total) local surface mass density, $\Sigma_{0} \geq
60 M_\odot\,{\rm pc}^{-2}$; (v) light bar, $M_B \simeq 2.0
\times 10^{10} M_\odot$; (vi) a rotation curve that falls to a
small asymptotic value, $v(50 \,{\rm kpc}) \mathrel{\mathpalette\fun <} 180\,{\rm km\,s^{-1}}$; and (vii) a
local escape velocity that is less than $420\,{\rm km\,s^{-1}}$.
What then are the prospects for falsifying the all-MACHO halo
hypothesis? Because $\tau_{\rm LMC}$ is apparently so small, it
may be difficult to accumulate sufficient statistics over the
next few years to exclude the possibility that $\tau_{\rm LMC}$
is as large as $2\times 10^{-7}$. It may be more promising to
establish that $\tau_{\rm bulge}$ is greater than $2\times
10^{-6}$, due to the higher microlensing rate toward the bulge,
Or, other observations could establish that the mass of the
bulge is in excess of $2\times 10^{10}M_\odot$, which cannot be
tolerated in models where the halo is entirely comprised of
MACHOs.
Several characteristics of an all-MACHO halo involve parameters
of the galactic model and the galactic rotation curve.
Improvements here could be equally decisive. The study of the
proper motions of satellite galaxies will further constrain the rotation curve
at large distances, and the recent observation of a
dwarf galaxy at a galactocentric distance of $16\,{\rm kpc}$
\cite{sagit} presents yet another opportunity. Continued
efforts to deduce the local escape velocity might well rule out
all-MACHO scenarios. A more precise determination of the local
circular velocity and position would also help limit the range
of viable models. Precision measurements of the pulse arrival
times for the binary pulsar PSR 1913+16 are reaching the level
of precision where the effects of solar acceleration, which
depends upon both $r_0$ and $v_c$, can be accurately determined
\cite{taylortiming}.
Equally interesting is testing the hypothesis of a no-MACHO
halo. Measurements of the event duration and light-curve
distortions due to parallax effects could help discriminate
between MACHOs in the halo and disk and/or LMC. Likewise, the
distribution of events in the LMC provides an important test of
whether or not the lenses are part of the LMC. It is probably
more difficult to determine whether or not the lenses are in the
spheroid (as opposed to the halo).
\subsection{\bf Summary}
Microlensing has already proven its utility as a probe of
the structure of the Galaxy. Based upon the existing data---which
is likely to represent but a small fraction of what will be
available over the next few years---and the extensive modeling
discussed here important conclusions can already be drawn.
First and foremost, the MACHO fraction of the galactic halo in
most viable models of the Galaxy is small---between 0\% and 30\%
(see Fig.~1). The few models with a halo MACHO fraction of
greater than 60\% are characterized by a very light halo. When
additional reasonable constraints that define the minimal extent
of the halo (such as local escape velocity and proper motions of
satellite galaxies) are taken into account none of these models
remain viable. The apparent elimination of the promising
baryonic candidate for the dark matter halo of our own galaxy
further enhances the case for cold dark matter and provides
further impetus for the efforts to directly detect cold dark
matter particles (e.g., neutralinos and axions).
Second, it is not impossible that the halo of the Galaxy
contains no MACHOs. If the optical depth for microlensing
toward the LMC is at the low end of the credible range, say less
than about $0.5\times 10^{-7}$, the microlensing events seen
could be due to microlensing by objects in the disk and/or LMC.
Or, it could be that the lenses are not halo objects, but rather
exist in a more centrally condensed component of the Galaxy
(e.g., the spheroid).
Third, based upon our modeling we conclude that the plausible
range for the local density of dark halo material is between
$6\times 10^{-25}{\,{\rm g\,cm^{-3}}}$ and $13 \times 10^{-25}{\,{\rm g\,cm^{-3}}}$, most
which is not in the form of MACHOs (see Fig.~2). This estimate
is about a factor of two higher than previous estimates because
we have taken the flattening of the halo into account
\cite{apjlett}.
Fourth, it is not possible to account for the large microlensing
rate in the direction of the bulge with an axisymmetric bulge; a
bar of mass $(2-3)\times 10^{10}M_\odot$ is required to meet our
minimal constraint $\tau_{\rm bulge} \ge 2\times 10^{-6}$.
Finally, while we are not able to rule out an all-MACHO halo
with certainty, our modeling points to future measurements that
could be decisive. The very few models with very-MACHO halos
($f_B \ge 0.75$) that survive our minimal set of constraints
have distinctive features that make allow them to be falsified:
$\tau_{\rm LMC} \simeq 2\times 10^{-7}$; $\tau_{\rm bulge}\simeq
2 \times 10^{-6}$; $v_c \le 215\,{\rm km\,s^{-1}}$; $\Sigma_{0} \geq 60
M_\odot\,{\rm pc}^{-2}$; $M_B \simeq 2.0 \times 10^{10}
M_\odot$; $v_c(50 \,{\rm kpc}) \mathrel{\mathpalette\fun <} 180\,{\rm km\,s^{-1}}$; $v_{\rm esc}\mathrel{\mathpalette\fun <} 420\,{\rm km\,s^{-1}}$.
\section*{Acknowledgments}
We thank C. Alcock, K. Cudworth and D. Bennett for helpful conversations. This
work was
supported in part by the DOE (at Chicago and Fermilab) and the
NASA (at Fermilab through grant NAG 5-2788).
|
\section*{Abstract}
This review of CP violation focuses on the status of the subject and its likely future development through experiments
in the Kaon system and with B-decays. Although present observations of CP violation are
perfectly consistent with the CKM model, we discuss the theoretical
and experimental difficulties which must be faced to establish this
conclusively. In so doing, theoretical predictions and experimental
prospects for detecting $\Delta S=1$ CP violation through
measurements of $\epsilon^\prime/\epsilon$ and of rare K decays are
reviewed. The crucial role that B CP-violating experiments will play in elucidating this issue is emphasized . The importance of looking for evidence for non-CKM
CP-violating phases, through a search for a non-vanishing transverse
muon polarization in $K_{\mu 3}$ decays, is also stressed.
\section{Introduction}
The discovery more than 30 years ago of the decay $K_{\rm L}\rightarrow 2\pi$
by Christianson, Cronin, Fitch and Turlay \cite{CCFT} provided the
first indication that CP, like parity, was also not a good symmetry
of nature. It is rather surprising that, for such a mature subject, we have
still so little experimental information available. Indeed, the only firm evidence for CP
violation to date remains that deduced from meausurements in the neutral Kaon system.
Here there are five parameters meausured: the values of the the two complex amplitude
ratios for the decays of $K_{\rm L}$ and $K_{\rm S}$ to $\pi^+\pi^-$ \newline ($n_{+-}=\epsilon+\epsilon^\prime$) and to $\pi^o\pi^o$ ($\eta_{oo}=\epsilon-
2\epsilon^\prime$), plus the semileptonic asymmetry in $K_{\rm L}$ decays ($A_{\rm K_L}$) \footnote{CP Lear \cite{Lear} has presented recently a preliminary meausurement of the
$K_{\rm S}$ semileptonic asymmetry, $A_{\rm K_S}$, which agrees with $A_{\rm K_L}$ within 10\%.}. However, in fact, these five numbers at present give only one
{\bf{independent}} piece of dynamical information. To a very good approximation
\cite{PDG}, $\eta_{+-}$ and $\eta_{oo}$ are equal in magnitude
\[
|\eta_{+-}|\simeq |\eta_{oo}| \simeq 2\times 10^{-3}~,
\]
and in phase
\[
\phi_{+-}\simeq \phi_{oo} \simeq 44^{\circ}~,
\]
with the ratio
\[
\epsilon^\prime/\epsilon \leq 10^{-3}~.
\]
Because the two pion intermediate states dominate the neutral Kaon width, assuming
CPT conservation \cite{DP} the phases $\phi_{+-}$ and $\phi_{oo}$ are
approximately equal to the superweak phase
\[
\phi_{\rm SW} = \tan^{-1}
\frac{2\Delta m}{\Gamma_{\rm S}-\Gamma_{\rm L}}=
(43.64\pm 0.14)^\circ~.
\]
CPT conservation also fixes the value of the semileptonic asymmetry in terms of
$ {\rm Re}~ \epsilon$ \cite{Cronin}. Since $\epsilon^{\prime}$ is so small, effectively one has \cite{DP}
\[
A_{\rm K_L} \simeq 2Re~ \eta_{+-}.
\]
Thus the dynamical information we have today is, essentially, that obtained in the
original discovery experiment\cite {CCFT}, augmented by the statement that there is
little or no $\Delta S=1$ CP violation!
The above statement is a bit of an exaggeration, since in the last 30 years we have
learned very much more about CP violation {\bf{outside}} the neutral Kaon complex. In particular,
very strong bounds have been established for the electric dipole moments
of the neutron and the electron \cite{PDG}
\[
d_e,d_n \leq 10^{-25}~{\rm ecm}~.
\]
Furthermore, we have uncovered the fundamental role that CP violation plays in the Universe,
to help establish the observed matter-antimatter asymmetry \cite{Sakharov}. In
addition, we have a variety of bounds on a host of other CP violating parameters, like
the amplitude ratios $\eta_{+-o}$ and $\eta_{ooo}$ for $K\rightarrow 3\pi$ decays , or
the transverse muon polarization $<P^{\mu}_{\perp}>$ in $K_{\mu 3}$ decays. These bounds, however, are
too insensitive to provide much dynamical information.
In the modern gauge theory paradigm
CP violation can have one of two possible origins. Either,
\begin{description}
\item[i)] the full Lagrangian of the theory is CP invariant, but this
symmetry is not preserved by the vacuum state: CP $|0\rangle \not=
|0\rangle$. In this case CP is a spontaneously broken
symmetry \cite{TDLee}.
\end{description}
Or
\begin{description}
\item[ii)] there are terms in the Lagrangian of the theory which
are not invariant under CP transformations. CP is explicitly broken
by these terms and is no longer a symmetry of the theory.
\end{description}
The first possibility, unfortunately, runs into a potential
cosmological problem\cite{KOZ}. As the universe cools below a
temperature $T^*$ where spontaneous CP violation occurs, one
expects that domains of different CP should form. These domains
are separated by walls having a typical surface energy density
$\sigma\sim T^{*^3}$. The energy density associated with these walls
dissipates slowly as the universe cools further and eventually
contributes an energy density to the universe at temperature T of order
$\rho_{\rm wall}\sim T^{*^3}T$. Such an energy density today would
typically exceed the universe closure density by many orders of
magnitude:
\[
\rho_{\rm wall}\sim 10^{-7}\left(\frac{T^*}{\rm TeV}\right)^3
{\rm GeV}^{-4} \gg \rho_{\rm closure} \sim 10^{-46}~{\rm GeV}^{-4}~.
\]
One can avoid this difficulty by imaging that the scale where CP is
spontaneously violated is very high, so that $T^*$ is above the
temperature where inflation occurs. In this case the problem
disappears, since the domains get inflated anyway. Nevertheless,
there are still problems since it proves difficult to connect
this high energy spontaneous breaking of CP with observed
phenomena at low energies. What emerges, in general, are models
which are quite complex \cite{Barr}, with CP violation being associated
with new interactions much as in the original superweak model of
Wolfenstein\cite{superweak}.
If, on the other hand, CP is explicitly broken the phenomenology of
neutral Kaon CP violation is a quite
natural result of the standard
model of the electroweak interactions. There is, however, a requirement
emerging from the demand of renormalizability which bears mentioning.
Namely, if CP is explicitly broken then renormalizability requires that
all the parameters in the Lagrangian which can be complex must be
so. A corollary of this observation
is that the number of possible CP violating
phases in a theory increases with the complexity of the theory, since there
are then more terms which can have imaginary coefficients.
In this respect, the three generation ($N_g=3$) standard model with
only one Higgs doublet is the simplest possible model, since it
has only one phase. With just one Higgs doublet,
the Hermiticity of the scalar potential allows
no complex parameters to appear. If CP
is not a symmetry, complex Yukawa couplings are, however, allowed.
After the breakdown of the $SU(2)\times U(1)$ symmetry, these couplings
produce complex mass matrices. Going to a physical basis with real
diagonal masses introduces a complex mixing matrix in the charged
currents of the theory. For the quark sector, this is the famous
Cabibbo-Kobayashi Maskawa (CKM) matrix\cite{CKM}.{\footnote{If the
neutrinos are massless, there is no corresponding matrix in the
lepton sector since it can be reabsorbed by redefining the neutrino
fields.}} This $N_g\times N_g$ unitary matrix contains $N_g(N_g-1)/2$
real angles and $N_g(N_g+1)/2$ phases. However, $2N_g-1$ of these phases
can be rotated away by redefinitions of the quark fields leaving only
$(N_g-1)(N_g-2)/2$ phases. Thus for $N_g=3$ the standard model has
only one physical complex phase to describe all CP violating
phenomena.{\footnote{This is not quite true. In the SM there is also another
phase related to the QCD vacuum angle which leads to a CP violating
interaction involving the gluonic field strength and its dual. We return to this point in the next section.}}
If CP is broken explicitly, it follows by the renormalizability
corollary that any extensions of the SM will involve further CP
violating phases. For instance, if one has two Higgs doublets,
$\Phi_1$ and $\Phi_2$, then the Hermiticity of the scalar potential
no longer forbids the appearance of complex terms like
\[
V = \ldots \mu_{12}\Phi_1^\dagger\Phi_2+\mu_{12}^*\Phi_2^\dagger\Phi_1~.
\]
Indeed, if one did not include such terms
the presence of complex
Yukawa couplings would induce such terms at one loop.
\section{CP Violation in the Standard Model: Expectations and Challenges}
The gauge sector of the $SU(3)\times SU(2)\times U(1)$ Standard Model contains no
explicit phases since the gauge fields are in the adjoint representation, which is
real, leading to real gauge couplings. Nevertheless, the nontrivial nature of the
gauge theory vacuum \cite{CDG} introduces a phase structure ($\theta$ vacua \cite{JR})
which allows for the presence of effective CP-violating interactions, involving the
non Abelian gauge field strengths and their duals:
\[
{\cal{L}}_{\rm CP~viol.}=\theta_{\rm strong} \frac{\alpha_{\rm s }}{8\pi}
F_a^{\mu\nu}\tilde F_{a\mu\nu}~+~\theta_{\rm weak} \frac{\alpha_{2}}{8\pi}
W_a^{\mu\nu}\tilde W_{a\mu\nu}~.
\]
The weak vacuum angle $\theta_{\rm weak}$ is actually irrelevant since the electroweak
theory is chiral and through a chiral rotation this angle can be set to zero \cite{AA}. The phase angle $\theta_{\rm strong}$, on the other hand, is problematic.
First of all, what contributes physically is not $\theta_{\rm strong}$, since
this angle receives
additional contributions from the weak interaction sector as a result of the chiral rotations
that render the quark mass matrices diagonal. Thus, in reality, the CP-violating efffective
interaction is
\[
{\cal{L}}_{\rm CP~viol.}={\bar\theta}\frac{\alpha_{\rm s }}{8\pi}
F_a^{\mu\nu}\tilde F_{a\mu\nu},
\]
where ${\bar\theta}=\theta_{\rm strong} +{\rm Arg}~{\rm det~M}$ and ${\rm M}$ is the quark mass matrix. The presence of such an
interaction in the Standard Model gives rise to a large contribution to the neutron electric dipole moment. One has, approximately \cite{BC},
\[
d_n\simeq \frac{e}{M_n}(\frac{m_q}{M_n}){\bar{\theta}}~,
\]
with $m_q$ a typical light quark mass.
Thus, to respect the existing experimental bounds on
$d_n$ \cite{PDG}, ${\bar{\theta}}$ must be extremely small:
\[
\bar\theta \leq 10^{-9}-10^{-10}.
\]
Why this should be so is unclear and constitutes
the strong CP problem \cite{strongCP}.
There are three possible attitudes one can take regarding the strong CP problem:
\begin{description}
\item[i)] One can just ignore this problem altogether. Afterall, ${\bar{\theta}}$ is
yet another uncalculable
parameter in the Standard Model, no different say than the unexplained ratio
$m_e/m_t \sim 10^{-6}$. So why should one worry about this parameter explicitly?
\item[ii)] One can try to calculate ${\bar{\theta}}$ and thereby "explain" the size
of the neutron electric dipole moment. To do so, one must imagine that CP
is spontaneously broken, so that indeed ${\bar{\theta}}$ is a finite calculable
quantity. However, as was mentioned earlier, then one runs into the domain wall
problem. Models that avoid this problem and which, in principle, produce a tiny
calculable ${\bar{\theta}}$ exist. However, they are quite recondite \cite{Barr}
and the price one pays for solving the strong CP problem this way is to introduce considerable
hidden underlying structure beneath the Standard Model.
\item[iii)] One can try to dynamically remove ${\bar{\theta}}$ from the theory. This
is my favorite solution, which I suggested long ago with Helen Quinn \cite{PQ}. Quinn and I proposed solving the strong CP problem by imagining that the Standard Model has
an additional global chiral symmetry. The presence of this, so called, $U(1)_{PQ}$
symmetry allows one to rotate away ${\bar{\theta}}$, much as the chiral nature of the
$SU(2) \times U(1)$
electroweak theory allows one to rotate away $\theta_{weak}$. However, this solution
also requires that axions exist \cite{WeW} and these elusive particles have yet to be detected \cite{strongCP}! An alternative possibility along this vein is that the Standard Model has a
natural chiral symmetry built in, which removes ${\bar{\theta}}$ because $m_u=0$
\cite{mzero}.However, this solution appears unlikely, as $m_u=0$ is disfavored by current algebra
analyses \cite{Leutwyler}.
\end{description}
It is fair to say that there is no clear understanding of what to do about the
strong CP problem at the moment. My own view is that the existence of this
unresolved problem is something that should not be ignored. There is a message here
and it may simply be that we do not understand CP violation at all!
In the Standard Model, with one Higgs doublet and three generations of quarks and
leptons, besides the strong CP phase ${\bar{\theta}}$ there is only one
other
CP-violating angle in the theory. This is the combination of phases in the Yukawa
couplings of the quarks to the Higgs doublet which remains after all redefinition
of quark fields, leading to a diagonal mass matrix, are done. This weak CP-violating
angle appears as a phase in the CKM mixing matrix, $V$, which details the coupling of the quarks to the charged
$W$-bosons:
\[
{\cal{L}}_{CC}=g_2{\bar{u}_i}\gamma_{\mu}(1-\gamma_5)V_{ij}d_jW^{\mu}~+~h.c.~.
\]
However, one does not really know if the complex phase present in the CKM
matrix is responsible for the CP violating phenomena observed
in the neutral Kaon system. Indeed, one does not know either whether
there are further phases besides the CKM phase. Nevertheless, it is
remarkable that, as a result of the hierarchial structure of the
CKM matrix and of other dynamical circumstances, one can {\bf qualitatively}
explain all we know experimentally about CP violation today on the
basis of the CKM picture.
\subsection{Testing the CKM Paradigm}
In what follows, I make use of the CKM matrix in the
parametrization adopted by the PDG\cite{PDG} and expand the three
real angles in the manner suggested by Wolfenstein\cite{Wolf} in
powers of the sine of the Cabibbo angle $\lambda$. To order $\lambda^3$
one has then
\[
V=\left|
\begin{array}{ccc}
1-\frac{\lambda^2}{2} & \lambda & A\lambda^3(\rho-i\eta) \\
\lambda & 1-\frac{\lambda^2}{2} & A\lambda^2 \\
A\lambda^3(1-\rho-i\eta) & -A\lambda^2 & 1
\end{array}
\right|
\]
with $A,\rho$ and $\eta$ being parameters one needs to fix from
experiments---with $\eta \not= 0$ signalling CP violation.\footnote{It
is often convenient instead of using $\rho-i\eta$ to write this in terms
of a magnitude and phase: $\rho-i\eta=\sigma e^{-i\delta}$, with $\delta$
being the CP violating CKM phase.}
This matrix, with is hierarchical interfamily structure, naturally accounts for the
three principal pieces of independent information that we have today on CP violation.
As we discussed earlier, these are:
\begin{description}
\item[i)] The value of the mass mixing parameter $|\epsilon| \sim 10^{-3}$,
which characterizes the strength of the $K_{\rm L}$ to $K_{\rm S}$ amplitude
ratios.
\item[ii)] The small value of the $\epsilon^\prime$ parameter which typifies direct ($\Delta S=1$) CP violation, with the ratio
$\epsilon^\prime/\epsilon \leq 10^{-3}$.
\item[iii)] The very strong bounds on the electric dipole moments
of the neutron and the electron, which give $d_e,d_n \leq 10^{-25}~{\rm ecm}$.
\end{description}
One can ``understand" the above three facts quite simply within the Standard Model and
the CKM paradigm. In the model the parameter $\epsilon$
is determined by the ratio of the imaginary to the real part of the
box graph of Fig. 1a. It is easy to check that this ratio is of
order
\[
\epsilon \sim \lambda^4 \sin\delta \sim 10^{-3} \sin\delta~.
\]
That is, $\epsilon$ is naturally small because of the suppression of
interfamily mixing without requiring the CKM phase $\delta$ to be
small.
The explanation of why $\epsilon^\prime/ \epsilon$ is small is a bit more dynamical.
Basically, this ratio is suppressed both because of the $\Delta I=1/2$ rule and
because $\epsilon^{\prime}$ arises through the Penguin diagrams of Fig.1b. These
diagrams involve the emission of virtual gluons (or photons \footnote{The contribution
of the electroweak Penguin diagrams are not suppressed by the $\Delta I=1/2$ rule, but these diagrams are only of $O(\alpha)$, not $O(\alpha_{\rm s})$.}), which are Zweig
suppressed\cite{GW}. Typically this gives
\[
\frac{\epsilon^\prime}{\epsilon} \sim \frac{1}{20} \cdot
\left[\frac{\alpha_{\rm s}}{12\pi}\ln \frac{m_t^2}{m_c^2}\right]
\sim 10^{-3}~.
\]
Finally, in the CKM model the electric dipole moments are small since
the first nonvanishing contributions\cite{Shabalin} occur at three
loops, as shown in Fig. 1c, leading to the estimate\cite{edm}
\[
d_{\rm q,e} \sim em_{\rm q,e}
\left[\frac{\alpha^2 \alpha_{\rm s}}{\pi^3}\right]
\left[\frac{m_t^2 m_b^2}{M_{\rm W}^6}\right] \lambda^6 \sin
\delta \sim 10^{-32}~{\rm ecm}~.
\]
\begin{figure}[t!]
~\epsfig{file=DAfig1.eps,width=13.5cm,height=5cm}
\caption{Graphs contributing to $\epsilon$, $\epsilon^{\prime}$ and $d_{\rm q,e}$}
\end{figure}
One can, of course, use the precise value of $\epsilon$ measured
experimentally to determine an allowed region for the parameters
entering in the CKM matrix. Because of theoretical uncertainties
in evaluating the hadronic matrix element of the $\Delta S=2$
operator associated with the box graph of Fig. 1a, this parameter
space region is rather large. Further restrictions on the allowed
values of CKM parameters come from semileptonic B decays and from
$B_d-\bar{B}_{ d}$ mixing. Because the parameter $A$, related to
$V_{cb}$, is better known, it has become traditional to present the
result of these analyses as a plot in the $\rho-\eta$ plane. Fig. 2
shows the results of a recent analysis, done in collaboration with my
student, K. Wang\cite{PW}. The input parameters used, as well as the
range allowed for certain hadronic amplitudes and other CKM matrix
elements is detailed in Table 1
\begin{table}[h!]
\caption[]{
\label{RhoParam}
Parameters used in the $\rho-\eta$ analysis of \cite{PW}
}
\begin{eqnarray*}
\begin{array}{rclr}
|\epsilon| & = & (2.26 \pm 0.02)\times 10^{-3} &\mbox{~~~~~~~\cite{PDG}} \\
\Delta m_d & = & (0.496 \pm 0.032) ps^{-1} & \mbox{\cite{Forty}} \\
m_t & = & (174 \pm 10^{+13}_{-12})~{\rm GeV} & \mbox{\cite{CDF}} \\
|V_{cb}| & = & 0.0378 \pm 0.0026 & \mbox{\cite{Stone}} \\
|V_{ub}|/|V_{cb}| & = & 0.08 \pm 0.02 & \mbox{\cite{Stone}} \\
B_{\rm K} & = & 0.825 \pm 0.035 & \mbox{\cite{Sharpe}} \\
\sqrt{B_d}~f_{B_d} & = & (180 \pm 30)~{\rm MeV} &\mbox{\cite{Lattice}}
\end{array}
\end{eqnarray*}
\end{table}
The resulting $1\sigma$ allowed contour emerging from the overlap
of the three constraints coming from $\epsilon$, $B_{\rm d}-\bar{B}_{\rm d}$
mixing and the ratio of $|V_{ub}|/|V_{cb}|$, shown in Fig. 3, gives a roughly symmetric
region around $\rho=0$ within the ranges
\[
0.2 \leq \eta \leq 0.5~; \;\; -0.4 \leq \rho \leq 0.4~.
\]
\begin{figure}[t!]
~\epsfig{file=DAfig2.eps,width=13.5cm,height=5cm}
\caption{Constraints on the $(\rho, \, \eta)$ plot}
\end{figure}
As anticipated by our qualitative discussion this region implies
that the CKM phase $\delta$ is large ($\rho=0$ corresponds to
$\delta=\pi/2$). One should note, however, that this analysis
does not establish the CKM paradigm. Using only the B physics
constraints one sees that in Fig. 2 there is also an overlap
region for $\eta=0$, which gives $\rho=-0.33 \pm 0.08$ \cite{PW}.
So one can still imagine that $\epsilon$ is due to some other CP
violating interaction, as in the superweak model \cite{superweak},
with the CKM phase $\delta$ being very small. As Wang
and I \cite{PW} discussed, one may perhaps eliminate this possibility by improving the bounds on $B_{\rm s}-\bar{B}_{\rm s}$ mixing to $\Delta m_s \geq 10 ~{\rm ps}^{-1}$. Since the
present LEP bound from ALEPH is $\Delta m_s \geq 6~{\rm ps}^{-1}$ \cite{ALEPH},
this is not going to be easy. Much more promising, however, is to
try to establish the correctedness of the CKM paradigm by looking at further
tests of CP violation, both in the Kaon system and by meausuring CP-violating
asymmetries in B-decays.
In principle, one can obtain quantitative tests of the CKM model purely with
Kaon experiments. However, the needed experiments are very challenging,
either due to the high precision required or due to the rarity of the
processes to be studied. Furthermore, the analysis of these results
is also theoretically very difficult, since it requires
better estimates of hadronic matrix elements than what we have at
present. A good example of both of these challenges is provided by
$\epsilon^\prime/\epsilon$. The present data on this ratio is
inconclusive, with the result obtained at Fermilab \cite{E731}:
\[
{\rm Re}~\frac{\epsilon^\prime}{\epsilon}=
(7.4 \pm 5.2 \pm 2.9) \times 10^{-4} ~~~\mbox{[E731]}~,
\]
being consistent with zero within the error, while the result
of the NA31 experiment at CERN \cite{NA31} giving a non-zero value
to $3\sigma$:
\[
{\rm Re}~\frac{\epsilon^\prime}{\epsilon}=
(23.0 \pm 3.6 \pm 5.4) \times 10^{-4} ~~~\mbox{[NA31]}~.
\]
Theoretically, the predictions for $\epsilon^\prime/\epsilon$ are
dependent both on the value of the CKM matrix elements and, more
importantly, on an estimate of certain hadronic matrix elements.
\begin{figure}
~\epsfig{file=DAfig3.eps,height=5cm}
\caption{Allowed region in the $\rho \, - \, \eta$ plane. Also shown in the plot is a possible unitarity triangle.}
\end{figure}
\subsubsection{Prospects for meausuring $\frac{\epsilon^\prime}{\epsilon}$}
There has been considerable activity recently to try to narrow down the expectations
for $\epsilon^\prime/\epsilon$ in the Standard Model. To describe the theoretical
status here, I find it useful to make use of an approximate formula for this ratio
derived by Buras and Lautenbacher \cite{BL}. These authors express the real
part of this ratio as the sum of two terms \footnote{ Because the difference between
the $I=0$ and $I=2$ $\pi\pi$ phase shifts is also near $45^\circ$ \cite{Gasser}, to
a good approximation $\epsilon^\prime/\epsilon \simeq {\rm Re}~\epsilon^\prime
/\epsilon$.}
\[
{\rm Re}~\frac{\epsilon^\prime}{\epsilon} \simeq
3.6\times 10^{-3} A^2\eta
\left[B_6-0.175\left(\frac{m^2_t}{M_{\rm W}^2}\right)^{0.93}B_8\right]~.
\]
Here $B_6$ and $B_8$ are quantities related to the matrix elements
of the dominant gluonic and electroweak Penguin operators,
respectively. The electroweak Penguin contribution is suppressed
relative to the gluonic Penguin contribution by a factor of
$\alpha/\alpha_{\rm s}$. However, as remarked earlier, it does not suffer from the
$\Delta I=3/2$ suppression and so one gains back a factor of about 20.
Furthermore, as Flynn and Randall \cite{FR} first noted, the
contribution of these terms can become significant for large top
mass because it grows approximately as $m_t^2$.
The result of the CKM analysis presented earlier brackets $A^2 \eta$ in the range
\[
0.12 \leq A^2\eta \leq 0.31~.
\]
For $m_t=175~{\rm GeV}$ the square bracket in the Buras-Lautenbacher formula
reduces to
[$B_6-0.75B_8$]. Hence one can write the expectation from theory
for $\epsilon^\prime/\epsilon$ as
\[
4.3 \times 10^{-4}[B_6-0.75B_8] \leq
{\rm Re}~\frac{\epsilon^\prime}{\epsilon} \leq 11.2 \times 10^{-4}
[B_6-0.75B_8]~.
\]
Because the top mass is so large, the predicted value for
$\epsilon^\prime/\epsilon$ depends rather crucially on {\bf both} $B_6$
and $B_8$. These (normalized) matrix elements have been estimated
by both lattice \cite{Ciuchini} and $1/N$ \cite{N} calculations
to be equal to each other, with an individual error of $\pm 20\%$:
\[
B_6=B_8=1 \pm 0.20~.
\]
Thus, unfortunately, the combination entering in $\epsilon^\prime/
\epsilon$ is poorly known. It appears that the best
one can say theoretically
is that ${\rm Re}~\epsilon^\prime/\epsilon$ should be a ``few" times
$10^{-4}$, with a ``few" being difficult to pin down more precisely.
Theory, at any rate, seems to favor the E731 experimental result
over that of NA31.
Fortunately, we may learn something more in this area in the near future.
There are 3rd generation experiments in preparation
both at Fermilab (KTeV) and CERN (NA48). These experiments should
begin taking data in a year or so and are designed to reach statistical
and systematic accuracy for $\epsilon^\prime/\epsilon$ at the level
of $10^{-4}$. The Frascati $\Phi$ factory DAPHNE, which should begin
operations in 1997, in principle, can also provide interesting information
for $\epsilon^\prime/\epsilon$. At DAPHNE one will need
an integrated luminosity of $\int {\cal{L}}~dt = 10~{\rm fb}^{-1}$ to
arrive at a statistical sensitivity for $\epsilon^\prime/\epsilon$
at the level of $10^{-4}$. However, if this statistical sensitivity
is reached, the systematic uncertainties will be quite different than
those at KTeV and NA48, providing a very useful cross-check. It is
important to remark that, irrespective of detailed theoretical
predictions, the observation of a non-zero value for $\epsilon^\prime/
\epsilon$ at a significant level would
provide direct evidence for $\Delta S=1$ CP violation and would rule
out a superweak explanation for the observed CP violation in the
neutral K sector.
\subsection{Rare Kaon Decays}
There are alternatives to the $\epsilon^\prime/\epsilon$ measurement
which could reveal $\Delta S=1$ (direct) CP violation. However, these
alternatives involve daunting experiments\cite{RW},
which are probably out of reach
in the near term. Whether these experiments can (or will?)
eventually be carried out is an open question. Nevertheless, it seems worthwhile here
to try to outline some of the theoretical expectations for these meausurments.
\subsubsection{$K_{\rm S}$ decays}
CP Lear already, and DAPHNE soon, will enable a
more thorough study of $K_{\rm S}$ decays by more efficient
tagging. The decay $K_{\rm S} \rightarrow 3\pi^o$ is CP-violating, while
the $K_{\rm S}\rightarrow \pi^+\pi^-\pi^o$ mode has both CP-conserving
and CP-violating pieces. However, even in this case the CP-conserving piece is small and vanishes in the center of the Dalitz
plot. Hence one can extract information about CP violation from
$K_{\rm S}\rightarrow 3\pi$ decays. The analogue $K_{\rm S}/K_{\rm L}$
amplitude ratios to $\eta_{+-}$ and $\eta_{oo}$ for $K\rightarrow 3\pi$
have both $\Delta S=1$ and $\Delta S=2$ pieces:
\[
\eta_{ooo}=\epsilon+\epsilon^\prime_{ooo}~; \;\;
\eta_{+-o}=\epsilon+\epsilon^\prime_{+-o}~.
\]
However,
in contrast to what obtains in the $K\rightarrow 2\pi$ case, here
the $\Delta S=1$ pieces can be larger. Cheng \cite{Cheng} gives
estimates for $\epsilon^\prime_{+-o}/\epsilon$ and
$\epsilon^\prime_{ooo}/\epsilon$ of $O(10^{-2})$, while others
are more pessimistic \cite{pessimistic}. Even so, there does not
appear to be any realistic prospects in the near future
to probe for $\Delta S=1$
CP-violating amplitudes in $K_{\rm S}\rightarrow 3\pi$. For instance, at
DAPHNE even with an integrated luminosity of $10~{\rm fb}^{-1}$
one can only reach an accuracy for $\eta_{+-o}$ and $\eta_{ooo}$
of order $3\times 10^{-3}$, which is at the level of $\epsilon$ not
$\epsilon^\prime$.
\subsubsection{Asymmetries in charged K-decays}
CP violating effects involving charged Kaons can only be due to
$\Delta S=1$ transitions, since $K^+\leftrightarrow K^-$ $\Delta S=2$
mixing is forbidden by charge conservation. A typical CP-violating
effect in charged Kaon decays necessitates a comparison between
$K^+$ and $K^-$ processes. However, a CP-violating asymmetry between
these processes can occur only if there are at least two decay
amplitudes involved and these amplitudes have {\bf{both}} a relative weak
CP-violating phase and a relative strong rescattering phase between
each other. Thus the resulting asymmetry necessarily depends on
strong dynamics.
To appreciate this fact, imagine writing the
decay amplitude for $K^+$ decay to a final state $f^+$ as
\[
A(K^+\rightarrow f^+)=A_1~e^{i\delta_{\rm W_1}}e^{i\delta_{\rm S_1}}+
A_2~e^{i\delta_{\rm W_2}}e^{i\delta_{\rm S_2}}~.
\]
Then the corresponding amplitude for the decay $K^-\rightarrow f^-$
is
\[
A(K^-\rightarrow f^-)=A_1~e^{-i\delta_{\rm W_1}}e^{i\delta_{S_1}}+
A_2~e^{-i\delta_{\rm W_2}}e^{i\delta_{\rm S_2}}~.
\]
That is, the strong rescattering phases are the same but one complex
conjugates the weak amplitudes. From the above, one sees that the
rate asymmetry between these processes is
\begin{eqnarray*}
{\cal{A}}(f^+;f^-)& = & \frac{\Gamma(K^+\rightarrow f^+)-
\Gamma(K^+\rightarrow f^-)}{\Gamma(K^+\rightarrow f^+)+
\Gamma(K^-\rightarrow f^-)} \\
& = & \frac{2A_1A_2\sin(\delta_{\rm W_2}-\delta_{\rm W_1})
\sin(\delta_{\rm S_2}-\delta_{\rm S_1})}
{A_1^2+A_2^2+2A_1A_2\cos(\delta_{\rm W_2}-\delta_{\rm W_1})
\cos(\delta_{\rm S_2}-\delta_{\rm S_1})}~.
\end{eqnarray*}
Unfortunately,
detailed calculations in the standard CKM paradigm for rate
asymmetries and asymmetries in Dalitz plot parameters for various
charged Kaon decays give quite tiny predictions. This can be
qualitatively understood as follows. The ratio
$A_2 \sin(\delta_{\rm W_2}-
\delta_{\rm W_1})/A_1$ is typically that of a Penguin amplitude to
a weak decay amplitude and so is of order $\epsilon^\prime/\epsilon$.
Furthermore, because of the small phase space for
$K\rightarrow 3\pi$ decays, or because one is dealing with electromagnetic
rescattering in $K\rightarrow \pi\pi\gamma$, the rescattering
contribution suppress these asymmetries even more. Table 2 gives
typical predictions, contrasting them to the expected reach of
the Frascati $\Phi$ factory with $\int {\cal{L}}~dt=10~{\rm fb}^{-1}$.
For the $K\rightarrow 3\pi$ decays, Belkov {\it et al.} \cite{Belkov}
give numbers at least a factor of 10 above those given in Table 2.
However, these numbers are predicated on having very large rescattering
phases which do not appear to be realistic\cite{IMP}. One is lead to
conclude that, if the CKM paradigm is correct, it is unlikely that
one will see a CP-violating signal in charged Kaon decays.
\begin{table}
\caption{Predictions for Asymmetries in $K^\pm$ Decays}
\begin{center}
\begin{tabular}{|c|c|c|} \hline
Asymmetry & Prediction & $\Phi$ Factory Reach \\ \hline
${\cal{A}}(\pi^+\pi^+\pi^-;\pi^-\pi^-\pi^+)$ &
$5\times 10^{-8}~~~\mbox{\cite{Pettit}}$ & $3\times 10^{-5}$ \\
${\cal{A}}(\pi^+\pi^o\pi^o;\pi^-\pi^o\pi^o)$ &
$2\times 10^{-7}~~~\mbox{\cite{Pettit}}$ & $5\times 10^{-5}$ \\
${\cal{A}}_{\rm Dalitz}(\pi^+\pi^+\pi^-;\pi^-\pi^+\pi^+)$ &
$2\times 10^{-6}~~~\mbox{\cite{Pettit}}$ & $3\times 10^{-4}$ \\
${\cal{A}}_{\rm Dalitz}(\pi^+\pi^o\pi^o;\pi^-\pi^o\pi^o)$ &
$1\times 10^{-6}~~~\mbox{\cite{Pettit}}$ & $2\times 10^{-4}$ \\
${\cal{A}}(\pi^+\pi^o\gamma;\pi^-\pi^o\gamma)$ &
$10^{-5}~~~\mbox{\cite{HYC}}$ & $2\times 10^{-3}$ \\ \hline
\end{tabular}
\end{center}
\end{table}
\subsubsection{$K_{\rm L}\rightarrow \pi^o\ell^+\ell^-;~K_{\rm L}
\rightarrow \pi^o\nu \bar\nu$}
Perhaps more promising are decays of the $K_{\rm L}$ to $\pi^o$ plus
lepton pairs. If the lepton pair is charged, then the process has
a CP conserving piece in which the decay proceeds via a $2\gamma$
intermediate state. Although there was some initial
controversy \cite{Seghal}, the rate for the process $K_{\rm L}\rightarrow
\pi^o\ell^+\ell^-$ arising from the CP-conserving $2\gamma$ transitiion
is probably at, or below, the $10^{-12}$ level \cite{Dan}:
\[
B(K_{\rm L}\rightarrow \pi^o\ell^+\ell^-)_{\rm CP~cons.}=
(0.3-1.2)\times 10^{-12}.
\]
Thus this contribution is just a small correction to the dominant CP-violating
amplitude arising from an effective spin 1 virtual state,
$K_{\rm L}\rightarrow \pi^oJ^*$. Since $\pi^o J^*$ is CP even,
this part of the amplitude is CP-violating. It
has two distinct pieces \cite{Dib}: an
indirect contribution from the CP even piece ($\epsilon K_1$) in the
$K_{\rm L}$ state, and a direct $\Delta S=1$ CP-violating
piece coming from the $K_2$ part of $K_L$:
\[
A(K_{\rm L}\rightarrow \pi^o J^*)=\epsilon A(K_1\rightarrow \pi^o J^*)+
A(K_2\rightarrow \pi^o J^*)~.
\]
To isolate the interesting direct CP contribution in this process
requires understanding first the size of the indirect contribution.
The amplitude $A(K_1\rightarrow \pi^o J^*)$ could be determined
absolutely if one had a measurement of the process
$K_{\rm S}\rightarrow \pi^o\ell^+\ell^-$. Since this is not at
hand, at the moment one has to rely on various guesstimates.
These give the following range for the indirect CP-violating
branching ratio\cite{WW}
\[
B(K_{\rm L}\rightarrow \pi^o\ell^+\ell^-)^{\rm indirect}_{\rm CP~violating}=
(1.6-6)\times 10^{-12}~,
\]
where the smaller number is the estimate coming from chiral
perturbation theory, while the other comes from
relating $A(K_1\rightarrow \pi^o J^*)$ to the analogous amplitude
for charged K decays.
The calculation of the direct CP-violating contribution to the
process $K_{\rm L}\rightarrow \pi^o\ell^+\ell^-$, as a result of
electroweak Penguin and box contributions and their gluonic
corrections, is perhaps the one that is most reliably known. The
branching ratio obtained by Buras, Lautenbacher, Misiak and M\"unz in
their next to leading order calculation
of this process\cite{BLMM} is
\[
B(K_{\rm L}\rightarrow \pi^o\ell^+\ell^-)^{\rm direct}_{\rm CP-violating}=
(5\pm 2)\times 10^{-12}~,
\]
where the error arises mostly from the imperfect knowledge of the CKM
matrix.
Experimentally one has the following 90\% C.L. for the two
$K_{\rm L}\rightarrow \pi^o\ell^+\ell^-$ processes:
\begin{eqnarray*}
B(K_{\rm L}\rightarrow \pi^o\mu^+\mu^-) & < & 5.1\times 10^{-9} \\
B(K_{\rm L}\rightarrow \pi^o e^+e^-) & < & 1.8\times 10^{-9}
\end{eqnarray*}
The first limit comes from the E799 experiment at Fermilab\cite{Harris},
while the second limit combines the bounds obtained by the E845
experiment at Brookhaven\cite{Ohl} and the E799 Fermilab
experiment\cite{DHH}. Forthcoming experiments at KEK and Fermilab
should be able to improve these limits by at least an order of
magnitude\footnote{The goal of the KEK 162 experiment is to get to a BR of $O(10^{-10})$ for this mode,
while KTeV hopes to push this BR down to $5\times 10^{-11}$.}, if they
can control the dangerous background arising from the decay
$K_{\rm L}\rightarrow \gamma\gamma e^+e^-$\cite{Greenlee}. Even more
distant experiments in the future may actually reach the level expected
theoretically for the $K_{\rm L}\rightarrow \pi^o e^+e^-$
rate \cite{WINS}.
However, it will be difficult to unravel the direct CP-violating
contribution from the indirect CP-violating contribution, unless
the $K_{\rm S}\rightarrow \pi^oe^+e^-$ rate is also measured
simultaneously.
In this respect, the process $K_{\rm L}\rightarrow \pi^o\nu \bar\nu$
is very much cleaner. This process is purely CP-violating, since it
has no $2\gamma$ contribution. Furthermore, it has a tiny indirect CP
contribution, since this is of order $\epsilon$ times the already
small $K^+\rightarrow \pi^+\nu\bar\nu$ amplitude\cite{Littenberg}.
Next to leading QCD calculations for the direct rate have been
carried out by Buchalla and Buras\cite{BBB}, who give the following
approximate formula for the branching ratio for this process
\[
B(K_{\rm L}\rightarrow \pi^o\nu\bar\nu)=8.2\times 10^{-11}
A^4\eta^2\left(\frac{m_t}{M_{\rm W}}\right)^{2.3}~.
\]
This value is very far below the present 90\% C.L. obtained by
the E799 experiment at Fermilab\cite{Weaver}
\[
B(K_{\rm L}\rightarrow \pi^o\nu\bar\nu)<5.8\times 10^{-5}~.
\]
KTeV should be able to lower this bound substantially, perhaps
to the level of $10^{-8}$, but this still leaves a long way to
go!
\subsubsection{$K^+\rightarrow \pi^+\nu\bar\nu$}
The last process I would like to consider in this section is the charged Kaon analogue
to the process just discussed. Although the decay
$K^+\rightarrow \pi^+\nu\bar\nu$ is not CP violating, it is
sensitive to $|V_{\rm td}|^2 \simeq A^2\lambda^6[(1-\rho)^2+\eta^2]$
and so, indirectly, it is sensitive to the CP violating CKM
parameter $\eta$.
For the CP violating decay $K_{\rm L}\rightarrow \pi^o\nu\bar\nu$
the electroweak Penguin and box contributions are dominated by loops
containing top quarks. Here, because one is not looking at the
imaginary part, one cannot neglect altogether the contribution from charm
quarks. If one could do so, the branching ratio formula for
$K^+\rightarrow \pi^+\nu\bar\nu$ would be given by an analogous
formula to that for $K_{\rm L}\rightarrow \pi^o\nu\bar\nu$ but with
$\eta^2\rightarrow \eta^2+(1-\rho)^2$.
Because $m_t$ is large, the $K^+\rightarrow \pi^+\nu\bar\nu$
branching ratio is not extremely sensitive to the contribution of
the charm-quark loops \cite{Dibc}. Furthermore, when next to leading
QCD corrections are computed the sensitivity of the result to the
precise value of the charm-quark mass is reduced considerably \cite{BB2}.
Buras {\it et al.}\cite{waiting} give the following approximate
formula for the $K^+\rightarrow \pi^+\nu\bar\nu$ branching ratio
\[
B(K^+\rightarrow \pi^+\nu\bar\nu)=2\times 10^{-11}
A^4\left[\eta^2+\frac{2}{3}(\rho^e-\rho)^2+\frac{1}{3}
(\rho^\tau-\rho)^2\right]
\left(\frac{m_t}{M_{\rm W}}\right)^{2.3}~.
\]
In the above the parameters $\rho^e$ and $\rho^\tau$ differ from
unity because of the presence of the charm-quark contributions. Taking
$m_t=175~{\rm GeV}$ and $m_c(m_c)=1.30 \pm 0.05~{\rm GeV}$ \cite{GL},
Buras {\it et al.}\cite{waiting} find that $\rho^e$ and $\rho^\tau$
lie in the ranges
\[
1.42 \leq \rho^e \leq 1.55~; \;\; 1.27 \leq \rho^\tau \leq 1.38~.
\]
Using the range of $\eta$ and $\rho$ determined by the CKM
analysis discussed earlier gives about a 40\% uncertainty for the $K^+\rightarrow \pi^+\nu\bar\nu$
branching ratio, leading to the expectation
\[
B(K^+\rightarrow \pi^+\nu\bar\nu)=(1\pm 0.4)\times 10^{-10}~.
\]
This number is to be compared to the best present limit coming
from the E787 experiment at Brookhaven. Careful cuts must be made
in the accepted $\pi^+$ range and momentum to avoid
potentially dangerous backgrounds, like $K^+\rightarrow \pi^+\pi^o$
and $K^+\rightarrow \mu^+\pi^o\nu$. There is
a new preliminary result for this branching
ratio \cite{LINS}
\[
B(K^+\rightarrow \pi^+\nu\bar\nu)<3\times 10^{-9}~~~
\mbox{(90\% C.L.)}
\]
which updates the previously published result from the E787
collaboration \cite{Atiyah}. This value is still about a factor
of 30 from the interesting CKM model range, but there are hopes
that one can get close to this sensitivity in the present run of
this experiment.
\subsection{The Promise of B-decay CP Violation}
If the CKM paradigm is correct, the analysis of current constraints on the CKM matrix shows that the CP-violating phase $\delta$ is sizable. The reason why one has {\bf small} CP-violating effects in the Kaon sector is solely due to the interfamily mixing suppression. In $B_d$ and $B_s$ decays one can involve all three generations directly and, in certain cases, one can obviate this supression altoghether \cite{BS}. Thus the study of CP violation in B-decays appears very promising.
To produce CP-violating effects, as usual, one has to have interference between two amplitudes that have different CP-violating phases. Because one is interested
in looking for potentially large sources of CP violation in B-decays, it is important to identify where in the B system sizable phases may reside. Within the CKM paradigm there are two places where large phases appear. The first of these is the relative phase between the $|B_d>$ and $|\bar{B}_d>$ states, which make up the neutral B mass eigenstates:
\[
|B_{d\pm}> \simeq \frac{1}{\sqrt 2}[(1+\epsilon_{B_d})|B_d> \pm (1-\epsilon_{B_d})|\bar{B}_d>]~.
\]
This $ B_d-\bar{B}_d$ mixing phase arises from a box-graph similar to that of Fig. 1a. Here, however, this graph is dominated by the contribution of the top quark loop and one has that
\[
\frac { (1-\epsilon_{B_d})}{ (1+\epsilon_{B_d})} \simeq \frac{V_{td}}{V_{td}^*} =e^{-2i\beta}~,
\]
where $\beta$ is the phase of the $td$ matrix element of the CKM matrix ($V_{td}=|V_{td}|e^{-i\beta}$). If the CKM phase $\delta$ is large so is, in general, $\beta$ since
\[
\tan~\beta=\frac{\sigma \sin~\delta}{1-\sigma \cos~\delta}=\frac{\eta}{1-\rho}~.
\]
The second place where a large phase appears is in any process involving, at the
quark level, a $b \rightarrow u$ transition since $V_{ub}=|V_{ub}|e^{-i\delta}$
provides precisely a meausure of the CKM phase.
The potentially large $B_d-\bar{B}_d$ CP-violating phase $\beta$ is a
prediction of the CKM paradigm which can be well tested, since this phase is
unpolluted by strong interaction effects. This is also the case for the
corresponding mixing phase arising from $B_s-\bar{B}_s$ mixing. However, in
this case the CKM prediction is that this phase should vanish, since $V_{ts}$
is approximately real. So for the case of $B_s$ decays the relevant tests are
null tests. At any rate, because of $B_d-\bar{B}_d$ mixing, a state
$|B_{d~{\rm phys}}(t)>$ which at $t=0$ was a pure $|B_d>$
state, evolves in time into a superposition of $|B_d>$ and $|\bar{B}_d> $
states:
\[
|B_{d~{\rm phys}}(t)>=e^{-im_{B_d}t}e^{-\Gamma_{B_d}t/2}[\cos\Delta m_{d}t/2~|B_d>
+ie^{-2i\beta}\sin\Delta m_{d}t/2~|\bar{B}_d>]~.
\]
A similar formula applies for $B_s$ decays.
It is also possible to isolate cleanly the CP-violating phases
appearing at the quark level by comparing the decays of $B$ mesons into some
definite final state $f$ to the corresponding
transition of $\bar{B}$ mesons to the charged-conjugate final state $\bar{f}$. If these transitions are dominated by a
single quark decay amplitude \cite{Rosner}, so that
\[
A( B \rightarrow f)=a_fe^{i\delta_f};~~~~~~~A( \bar{B} \rightarrow
\bar{f})=a_f^*e^{i\delta_f}~,
\]
where $\delta_f$ is a strong rescattering phase, then the ratio of these two
amplitudes will be directly sensitive to the quark decay phase. In these
circumstances, for decays involving $b\rightarrow u $ transitions, the ratio
\[
\frac{A( \bar{B} \rightarrow \bar{f})}{ A( B \rightarrow f)}\simeq
\frac{A(b\rightarrow uq\bar{q'})}{A(\bar{b}\rightarrow \bar{u}\bar{q}q')}=
\frac{V_{ub}}{V_{ub}^*}=e^{-2i\delta}
\]
is a measure of the CP-violating phase $\delta$. On the other hand, the
corresponding ratio of amplitudes which involve a $b \rightarrow c$ transition at
the quark level will contain no large CP-violating phase at all, since $V_{cb}$ is real.
In view of the above considerations, the best way to study CP violation in
neutral $B$-decays is through a comparison of the time evolution of decays of
states "born" as a $B_d$ (or a $B_s$) into final states $f$, which are CP self-conjugate
[$\bar{f}=\pm f$], to the corresponding time evolution of states which were born as
a $\bar{B}_d$ (or a $\bar{B}_s$) and decay to $\bar{f}$ \cite{Bunk}.
A straightforward calculation gives for the time dependent rates for these
processes the expressions:
\[
\Gamma(B_{phys}(t)\rightarrow f)=\Gamma(B \rightarrow
f)e^{-\Gamma_Bt}[1-\eta_f\lambda_f\sin\Delta m_Bt]
\]
\[
\Gamma(\bar{B}_{phys}(t)\rightarrow \bar{f})=\Gamma(B \rightarrow
f)e^{-\Gamma_Bt}[1+\eta_f\lambda_f\sin\Delta m_Bt]
\]
In the above $\eta_f$ characterizes the CP parity of the state $f$,
with $\bar{f}=\eta_f f $ and $\eta_f=\pm 1$, while $\lambda_f$ encapsulates the
mixing and decay CP violation information for the process. For all decays
$B\rightarrow f$ which are dominated by just {\bf{one}} weak decay amplitude
the parameter $\lambda_f$ is free of strong interaction complications and
takes one of four values, depending on whether one is dealing with a $B_d$ or
$B_s$ decay and on whether the decay processes at the quark level involves a $
b\rightarrow c$ or a $b \rightarrow u$ transition. In these circumstances
$\lambda_f$ meausures purely CKM information and one finds
\[ \begin{array}{ll}
\lambda_f=\sin 2\beta & [B_d ~\rm{decays};~b \rightarrow c ~\rm{transition}] \\
\lambda_f=\sin 2(\beta +\delta)\equiv \sin 2\alpha & [B_d ~\rm{decays};~b \rightarrow u ~\rm{transition}] \\
\lambda_f=0 & [B_s ~\rm{decays};~b \rightarrow c ~\rm{transition}] \\
\lambda_f=\sin 2\delta \equiv \sin 2 \gamma & [B_s ~\rm{decays};~b
\rightarrow u ~ \rm{transition}]~.
\end{array} \]
The angles $\alpha, \beta$ and $\gamma$ entering in the above equations have
a very nice geometrical interpretation \cite{bj}. They are the angles of the,
so called, unitarity triangle in the $\rho-\eta$ plane, whose base is along
the $\rho$-axis going from $\rho=0$ to $\rho=1$ and whose apex is the point
$(\rho,\eta)$. That this is the case can be easily deduced by considering the
$bd$ matrix element of the CKM unitarity equation ($V^{\dagger}V=1$):
\[
V_{ub}^*V_{ud}+V_{cb}^*V_{cd}+V_{tb}^*V_{td}=0~.
\]
To leading order in $\lambda$, the above equation reduces to
\[
V_{ub}^*+V_{td}= A\lambda^3
\]
which, upon dividing by $A\lambda^3$, is precisely the equation describing the
unitarity triangle. One possible unitarity triangle, with the angles
$\alpha,\beta$ and $\gamma$ identified, is shown in Fig. 3.
Because our present knowledge of the CKM matrix still allows a considerable
range for $\rho$ and $\eta$, as shown in Fig. 3, there is considerable
uncertainty on what to expect for $\sin 2\alpha$, $\sin 2\beta$ and $ \sin 2
\gamma$. Nevertheless, from an analysis of the allowed region in the
$\rho-\eta$ plane, one can infer the allowed ranges for the unitarity triangle
angles. In general, one finds that while both $ \sin 2\alpha$ and $\sin
2\gamma$ can vanish, $\sin 2 \beta$ is both {\bf{nonvanishing}} and
{\bf{large}} \cite{Nir}. This is illustrated in Fig. 4, taken from my recent
examination of this question with Wang \cite{PW}, which plots the presently
allowed region in the $\sin 2\alpha-\sin 2\beta$ plane. One sees from this
figure that
\[
0.23\leq \sin 2 \beta \leq 0.84~.
\]
Thus, in contrast to CP violation phenomena in the Kaon system, for the B
system within the CKM paradigm there are places where one expects to see large
effects.
From the above discussion, it is clear that a particularly clean test of the
CKM paradigm would be provided by the meausurement of $\sin 2\beta$, via the
observation of a difference in the rates of specific $B_d$ and $\bar{B}_d$
decays to CP self-conjugate states which involve a $b \rightarrow c$
transition. A favored mode to study is the decay $B_d \rightarrow \psi K_S$
along with its conjugate \cite{NQ}. This decay has a largish branching ratio \cite{PDG}
and quite a distinct signature from the leptonic decay of the $\psi$.
Furthermore, one can argue that this decay is quite clean theoretically.
Recall that the identification of $\lambda_f$ with one of the angles of the
unitarity triangle required that the decay rate for the process $B \rightarrow
f$ be dominated by a single decay amplitude. This, in general, is only
approximately true. For instance, for the process in question, at the quark
level both a $ b \rightarrow c$ decay graph {\bf{and}} a $b \rightarrow s$
Penguin graph contribute. However, although this decay involves more than one
amplitude, both of these amplitudes have the {\bf{same}} weak decay phase
\cite{GLP}. The amplitude involving the quark decay graph has no weak phase
since it involves a $ b \rightarrow c$ transition. This is also true for the
$b \rightarrow s$ Penguin graph, since this graph is dominated by the top loop
contribution which is dominantly real. Hence, effectively, the ratio of
$A(\bar{B}_d \rightarrow \psi K_S)$ to $A(B_d \rightarrow \psi K_S)$ is, to a
very good approximation, unity and $\lambda_{\psi K_S}$ indeed meausures $\sin
2 \beta$.
Even though $\sin 2\beta$, at least in the CKM paradigm, is large, the
meausurement of $\lambda_{\psi K_S}$ is far from trivial, since one must be
able to determine whether the decaying state was born as a $B_d$ or a
$\bar{B}_d$ and one must have sufficient rate to detect the produced $\psi$
though its small leptonic decay mode. Nevertheless, it is clear that future
meausurements of this and allied decay modes (like $B_d\rightarrow \psi K^*
\rightarrow \psi K_S \pi $ \cite{KKPS}) perhaps at HERA-B, but certainly at the B factories
under construction at SLAC and KEK and eventually in hadron colliders, offers
an excellent chance of verifying- or put into question- the CKM paradigm.
The prospects of meausuring the other two angles of the unitarity triangle,
$\alpha$ and $\gamma$, and thus of checking the CKM prediction $\alpha + \beta
+\gamma=\pi$, appear more difficult. These meausurments are, nevertheless,
quite important. As Winstein \cite{WEIN} has pointed out, even if one were to
meausure a large value for $\sin 2\beta$ this still does not totally exclude
a superweak explanation. One could imagine a perverse superweak model which,
somehow, contained a small $\Delta S=2$ CP-violating mixing phase, of
$O(\epsilon)$, but a large $\Delta B=2$ CP-violating mixing phase, $2\beta$.
This model can be ruled out by meausuring the angle $\alpha$ independently
since, as there are no decay phases, it predicts that $\sin 2\beta=\sin
2\alpha$. As Fig. 4 shows, this "superweak' prediction is not excluded by
present day data. Thus, if by chance, future meausurements of $\alpha$ and
$\beta$ were to fall on the superweak line one would still need a measurement
of $\gamma$ to settle the issue \footnote{In superweak models the angle
$2\gamma $ is the phase associated with $ B_s-\bar{B}_s$ mixing and does not
necessarily vanish. However, in these models all $B_s$ decays to CP
self-conjugate states, irrespective of whether they involve a $b \rightarrow
u$ or a $b \rightarrow c$ transition, should produce $\lambda_f=\gamma$.}.
\begin{figure}
~\epsfig{file=DAfig4.eps,width=12.5cm,height=5cm}
\caption[]{ Plot of the allowed region in the $\sin 2 \alpha -\sin 2 \beta$
plane. The vertical line corresponds to the superweak prediction
$\sin 2 \alpha =\sin 2 \beta$ \cite{WEIN} .}
\end{figure}
Most likely, the decay mode $B_d \rightarrow \pi^+\pi^-$ is the process that
is best suited to study the angle $\alpha$ \cite{NQ}. However, the branching ratio for
this mode is not yet totally in hand, but is probably quite small, of
$O(10^{-5})$ \footnote{ Recently, CLEO has observed a few $B_d$ decays into
pairs of light mesons and has been able to determine the branching ratio for
the sum of the decay rates into both $\pi K$ and $\pi \pi$ final states:
$BR(B_d \rightarrow \pi K+\pi\pi) =(1.8 \pm 0.6 \pm 0.2)\times 10^{-5}$ \cite{CLEO}.}.
This mode also may suffer some Penguin pollution, with estimates ranging from
1 \% to 10 \% in the amplitude \cite{GLP}. The CLEO \cite{CLEO} indications
that the branching ratio of $B_d\rightarrow \pi K$ is approximately the same
as that for $B_d \rightarrow \pi \pi$ give one already some assurances that
the Penguin amplitude in $B_d \rightarrow \pi^+\pi^-$ cannot dominate the
process, since this amplitude is smaller by a factor of $|V_{td}|/|V_{ts}|$
compared to the $B_d\rightarrow \pi K$ amplitude. Furthermore, in principle,
one can isolate the contribution of the Penguin graphs in the process $B_d
\rightarrow \pi^+\pi^-$
by meausuring in addition the decays $B_d \rightarrow \pi^o \pi^o$ and
$B^+\rightarrow \pi^+\pi^o$ \cite{GL2}. These extra meausurements allow for a
complete isospin analysis of the amplitudes entering in these two body decays
and an isolation of the relative rescattering phases. Similar techniques
\cite{Quinn} may allow the extraction of the angle $\alpha$ in other decay
modes, like $B_d \rightarrow \rho\pi$, which are not CP self-conjugate.
The determination of the angle $\gamma$ is even more problematic. In
principle, this angle can be determined by studying the time evolution of
certain $B_s$ decays. Here, perhaps, the best mode to study would be the decay
$B_s \rightarrow \pi^o K_S$. However, at the asymmetric B factories now
under construction, meausurements of $B_s$ decays are unlikely. These
colliders are optimized to operate at the $\Upsilon(4s)$ and running above the
$B_s$ production threshold will entail substantial loss of luminosity. Thus a
determination of $\sin 2\gamma$ from $B_s$ decays will have to await
dedicated experiments at hadron colliders. One may, however, be able to
determine $\gamma$, and thus the CKM phase $\delta$, before that by utilizing
different techniques. For instance, Gronau and Wyler \cite{GW2}, have shown
that one could, in principle, extract $\gamma$ by studying the charged B
decays $B^{\pm}\rightarrow D K^{\pm}$ and their neutral counterparts, using
isospinology to isolate the rescattering phases. It remains to be seen,
however, whether this approach can really bear fruit in the presence of
experimental errors.
\section{Looking for new CP-violating phases}
I would like to argue now a little more broadly about tests of CP violation.
Obviously, it is
very important to check whether the CKM paradigm is correct. Positive signals
for
$\epsilon^\prime/\epsilon \not= 0$ will indicate
the general validity of the CKM picture, since they require the presence of a
$ \Delta S=1$
phase. However, given the large
theoretical uncertainty on the value of this quantity, it is clear that values
of
$\epsilon^\prime/\epsilon$ consistent with zero at the $10^{-4}$ level cannot
disprove
this picture. In my view, it is likely that what will
provide the crucial smoking gun for the CKM paradigm are searches
for CP violation in the B system. The detection of the expected large
asymmetry in $B_d\rightarrow \psi K_{\rm S}$ decays is of paramount
importance,
with the meausurement of the other angles in the unitarity triangle and of
rare Kaon
decays providing eventually a more detailed picture. However, whether the CKM
picture is (essentially) correct or not, it is also importat to mount
experiments
which may provide the first glimpse at {\bf other} CP-violating
phases, besides the CKM phase $\delta$. Of course, if the CKM picture is
incorrect then
one knows that at least some of these new phases are superweak in nature,
arising
in the $\Delta S=2$ (and, perhaps, in the $\Delta B=2$) sectors. Even if the
CKM paradigm
is essentially correct, there may be other phases which produce small
violations in the
fermion mixing matrix but which are important elsewhere.
Indeed, there are good theoretical arguments for having further CP-violating
phases, besides the CKM phase $\delta$. For instance, to establish
a matter-antimatter asymmetry in the Universe one needs to have
processes which involve CP violation\cite{Sakharov}. If the origin
of this asymmetry comes from processes at the GUT scale, then, in
general, the GUT interactions contain further CP-violating phases besides
the CKM phase $\delta~$\cite{PecceiB}. If this asymmetry is
established at the electroweak scale\cite{Shap}, then most likely
one again needs further phases, both because intrafamily suppression
gives not enough CP violation in the CKM case to generate the asymmetry
and because one needs to
have more than one Higgs doublet\cite{Cohen}. Indeed this last point
gives the fundamental reason why one should expect to have further
CP-violating phases, besides the CKM phase $\delta$. It is likely
that the standard model is part of a larger theory. For instance,
supersymmetric extensions of the SM have been much in vogue
lately. Any such extensions will introduce further particles and
couplings and thus, by the simple corollary mentioned in the Introduction,
they will introduce new CP-violating phases.
The best place to look for non-CKM phases is in processes where
CP violation within the CKM paradigm is either vanishing or very
suppressed. One such example is provided by experiments aimed at
measuring the electric dipole moments of the neutron or the electron,
since electric dipole moments are predicted to be extremely small
in the CKM model. Another example concerns measurements of the
transverse muon polarization $\langle P_\perp^\mu\rangle$ in
$K_{\mu 3}$ decays, which vanishes in the CKM paradigm\cite{Leurer}.
The transverse muon polarization measures a T-violating triple
correlation\cite{Sakurai}
\[
\langle P_\perp^\mu\rangle \sim \langle
\vec s_\mu \cdot (\vec p_\mu\times \vec p_\pi)\rangle~.
\]
In as much as one can produce such an effect also as a result of
final state interactions (FSI) this is not a totally clean test
for new CP-violating phases. With two charged particles in the
final state, like for the decay $K_{\rm L}\rightarrow \pi^-\mu^+\nu_\mu$,
one expects FSI to give typically
$\langle P_\perp^\mu\rangle_{\rm FSI} \sim \alpha/\pi \sim
10^{-3}$~\cite{Adkins}. However, for the process
$K^+\rightarrow \pi^o\mu^+\nu_\mu$ with only one charged particle
in the final state, the FSI effects should be much smaller. Indeed,
Zhitnitski\cite{Zhitnitski} estimates for this proceses that
$\langle P_\perp^\mu\rangle_{\rm FSI}\sim 10^{-6}$. So
a $\langle P_\perp^\mu\rangle$ measurement in the $K^+\rightarrow
\pi^o\mu^+\nu_\mu$ decay is a good place to test for additional
CP-violating phases.
The transverse muon polarization $\langle P_\perp^\mu\rangle$ is
particularly sensitive to scalar interactions and thus is a good
probe of CP-violating phases arising from the Higgs
sector\cite{CFK}. One can write the effective $K_{\mu 3}$
amplitude\cite{BG} as
\[
A=G_{\rm F}\sin\theta_c f_+(q^2)
\left\{p_K^\mu \bar\mu \gamma_\mu(1-\gamma_5)
\nu_\mu + f_S(q^2)m_\mu \bar\mu(1-\gamma_5)\nu_\mu\right\}~.
\]
Then
\[
\langle P_\perp^\mu\rangle = \frac{m_\mu}{M_{\rm K}} {\rm Im}~f_{\rm S}
\left[\frac{|\vec p_\mu|}
{E_\mu+|\vec p_\mu|n_\mu \cdot n_\nu-m_\mu^2/M_{\rm K}}\right]
\simeq 0.2~{\rm Im}~f_{\rm S}~.
\]
Here $n_\nu(n_\nu)$ are unit vectors along the muon (neutrino)
directions and the numerical value represents the expectation after
doing an average over phase space\cite{Kuno}.
\begin{figure}
~\epsfig{file=DAfig5.eps,width=13.5cm,height=3.0cm}
\caption{Graphs contributing to $\langle P_\perp^\mu\rangle$}
\end{figure}
Contributions to ${\rm Im}~f_{\rm S}$ can arise in multi-Higgs models
(like the Weinberg 3-Higgs model\cite{Weinberg}) from the charged
Higgs exchange shown in Fig. 5, leading to \cite{Higgs}
\[
{\rm Im}~f_{\rm S} \simeq {\rm Im}(\alpha^*\gamma)
\frac{M_{\rm K}^2}{M_{{\rm H}^-}^2}~.
\]
Here $\alpha(\gamma)$ are constants associated with the charged
Higgs coupling to quarks (leptons). Because a leptonic vertex
is involved, one in general does not have a strong constraint on
${\rm Im}(\alpha^*\gamma)$. By examining possible non-standard
contributions to the B semileptonic decay $B\rightarrow X\tau \nu_\tau$,
Grossman\cite{Grossman} obtains
\[
{\rm Im}(\alpha^*\gamma) <
\frac{0.23~M^2_{H^-}}{[{\rm GeV}]^2}
\]
which yields a bound for $\langle P_\perp^\mu\rangle$ of
$\langle P_\perp^\mu\rangle < 10^{-2}$. Amusingly, this is
the same bound one infers from the most accurate measurement of
$\langle P_\perp^\mu\rangle$ done at Brookhaven about a decade
ago \cite{Blatt}, which yielded
\[
\langle P_\perp^\mu\rangle = (-3.1\pm 5.3)\times 10^{-3}~.\]
In specific models, however, the leptonic phases associated with
charged Higgs couplings are correlated with the hadronic phases.
In this case,
one can obtain more specific restrictions on $\langle P_\perp^\mu\rangle$
due to the strong bounds on the neutron electric dipole moment.
For instance, for the Weinberg 3 Higgs model, one relates
${\rm Im} (\alpha^*\gamma)$ to a similar product of couplings of
the charged Higgs to quarks\cite{Higgs}:
\[
{\rm Im}(\alpha^*\gamma)=\left(\frac{v_u}{v_e}\right)^2~
{\rm Im}(\alpha^*\beta)~,
\]
where $v_u~(v_e)$ are the VEV of the Higgs doublets which couples
to up-like quarks (leptons). The strong bound on the neutron electric
dipole moment\cite{PDG} then gives the constraint
\[
{\rm Im}(\alpha^*\beta) \leq
\frac{4\times 10^{-3}~M_{{\rm H}^-}^2}{[{\rm GeV}]^2}~.
\]
If one assumes that $v_u \sim v_e$, this latter bound gives
a strong constraint on $\langle P_\perp^\mu\rangle \break
[\langle P_\perp^\mu\rangle < 10^{-4}]$.
However, this constraint is removed if
$v_u/v_e \sim m_t/m_\tau$.
Similar results are obtained in the simplest supersymmetric
extension of the SM. In this case, ${\rm Im}~f_{\rm S}$ arises from
a complex phase associated with the gluino mass. Assuming all
supersymmetric masses are of the same order, Christova and
Fabbrichesi\cite{CF} arrive at the estimate
\[
{\rm Im}~f_{\rm S} \simeq \frac{M_{\rm K}^2}{m_{\tilde g}^2}
\frac{\alpha_{\rm s}}{12\pi} \sin\phi_{\rm susy}~,
\]
where $\phi_{\rm susy}$ is the gluino mass CP-violating phase.
This phase, however, is strongly restricted by the neutron electric
dipole moment. Typically, one finds\cite{Hall}
\[
\sin\phi_{\rm susy} \leq \frac{10^{-7}~m^2_{\tilde g}}
{[{\rm GeV}]^2}
\]
leading to a negligible contribution for
$\langle P_\perp^\mu\rangle$, below the level of
$\langle P_\perp^\mu\rangle_{\rm FSI}$.
An experiment (E246) is presently underway at KEK aimed at
improving the bound on $\langle P_\perp^\mu\rangle$ obtained
earlier at Brookhaven. The sensitivity of E246 is such that one should
be able to achieve an error $\delta\langle P_\perp^\mu\rangle \sim
5\times 10^{-4}$\cite{Kuno}. This level of precision is very
interesting and, in some ways, it is comparable or better to
$d_n$ measurements for probing CP-violating phases from the scalar
sector. This is the case, for instance, in the Weinberg model if
$v_u/v_e$ is large. At any rate, if a positive signal were
to be found, it would be a clear indication for a non-CKM CP-violating
phase. Furthermore, as Garisto\cite{Garisto} has pointed out, a
positive signal at the level aimed by the E246 experiment would
imply very large effects in the corresponding decays in the B system
involving $\tau$-leptons (processes like $B^+\rightarrow D^o \tau^+
\nu_\tau$), since one expects, roughly,
\[
\langle P_\perp^\tau\rangle_{\rm B} \sim
\frac{M_{\rm B}}{M_{\rm K}} \frac{m_\tau}{m_\mu}
\langle P_\perp^\mu\rangle_{\rm K}~.
\]
Thus, in principle, a very interesting experimental cross-check
could be done.
\section{Concluding Remarks}
I would like to conclude more or less in the way in which I started this
review,
by reemphasizing that even after thirty years from the discovery of CP
violation this
phenomena remains shrouded in mystery. However, there are some grounds for
optimism.
It is quite possible that before the year 2000 we shall know whether the CKM
model
provides the approximately correct description of CP violation. For instance,
a convincing
non zero determination of $\epsilon^{\prime}/\epsilon$ would exclude the
superweak
hypothesis, while a meausurement of $\sin 2\beta \sim O(1)$ would strongly
favor the
CKM explanation. Both of these experimental results could be on hand in this
time frame.
A more detailed understanding of the full CKM structure, or a further
understanding of
CP violation if the CKM paradigm fails, will be more difficult. The
meausurement of
$\alpha$ is probably the simplest of the more difficult things to accomplish.
A
direct meausurement of the CKM phase $\delta$, or equivalently the angle
$\gamma$, by
means of experiments in the B sector or through the study of rare K decays,
is very
challenging. So are attempts at finding non CKM phases, although experiments
searching
for these effects are to be encouraged since they would signal new physics
beyond the Standard Model. Indeed, it is important also that other CP
violation
experiments where one expects very small effects in the Standard Model be
pushed
to their limits, as surprises may arise. A case in point is provided by
searches for CP violation in charged K decays, or in $K_S$ decays,
to be carried out here at DAPHNE, where little is really known. It is unclear,
however,
whether all this experimental activity will be able to throw any light on
the strong CP problem. Nevertheless, if we are to really understand CP
violation, one
day we will have to understand why $\bar{\theta}\simeq 0$.
\section*{Acknowledgments}
Part of the material on CP violation in K-decays presented here is drawn from a talk I gave at the 23rd INS Symposium in Tokyo, Japan \cite{PINS}. This work was supported in part by the
Department of Energy under Grant No. FG03-91ER40662.
\vspace{1pc}
|
\chapter{Introduction}
Recent years have witnessed a remarkably intense experimental and
theoretical activity in search of scale-invariance
and fractality in multihadron production processes,
for short also called ``intermittency''.
These investigations cover all types of reactions
ranging from e$^+$e$^-$ annihilation to nucleus-nucleus collisions,
up to the highest attainable energies.
The creation of soft hadrons in these processes, a major
fraction of the total cross section, relates to the
strong-coupling long-distance regime of Quantum Chromodynamics (QCD),
at present one of the least explored sectors in the whole of
high-energy particle physics.
A primary motivation is the expectation that
scale-invariance or self-similarity,
analogous to that
often encountered in complex non-linear systems, might open new
avenues ultimately leading towards deeper insight into long-distance
properties of QCD and the unsolved problem of colour confinement.
History shows that studies of fluctuations have
often triggered significant advances in physics.
In the present context, it was the observation of
``unusually large'' particle density fluctuations, reminiscent
of intermittency spikes in spatio-temporal turbulence, which prompted
the pioneering suggestion to investigate the pattern of
multiplicity fluctuations in multihadron events for ever decreasing
domains of phase-space.
Scale-invariance or fractality would manifest itself in
power-law behaviour for scaled factorial moments
of the multiplicity distribution in such domains.
It is important to stress here that, in practice, one deals with the
problem of evolution of particle number distributions for ever smaller
bins and intermittent behaviour implies that, for small phase space bins,
the distributions become wider in a specific way. The same problem can be
stated as an increasing role of correlations within a small phase space
volume.
Through a multitude of increasingly sophisticated
experimental studies of factorial moments,
much new information has been gathered in a surprisingly short time.
This work indeed confirms approximate power behaviour down to the
experimentally possible resolution,
especially when carried out in two- and three-dimensional phase space.
The proposal to look for intermittency also has triggered
a thorough revival of interest in the old subject of particle correlations,
by experimentalists and theorists alike. The need
for greater sensitivity in measurements of correlation
functions has directly inspired important work on refined
analysis techniques. A promising and long overdue systematic
approach to correlation phenomena of various sorts,
including Bose-Einstein interferometry, is finally emerging.
The large body of experimental observations
now available is calling for satisfactory explanation and, indeed,
theoretical ideas of all sorts abound.
The level of theoretical understanding is quite different
for the various types of collision processes.
In e$^+$e$^-$ annihilation, parton cascade models based on
leading-log QCD have met considerable success and
good overall description of multiplicity fluctuations is claimed.
Closer inspection, nevertheless, reveals potentially serious
deviations from the data, thus requiring further study.
For other processes, in particular hadron initiated collisions,
models are faced with large and partly unexpected obstacles.
This may be a reflection of
insufficient knowledge of the reaction dynamics, although
present evidence points to hadronization as the main culprit.
Within the framework of perturbative QCD, results
of considerable interest on the
emergence of power behaviour and multifractality have been obtained.
However, these are asymptotic in nature and most likely quite
unrelated to present-day experiment. Being
related to the mechanism of confinement, not surprisingly,
the role of hadronization remains unclear.
Random self-similar multiplicative branching models have inspired
much of the original work on intermittency.
Among many scale-invariant physical systems,
the cascade process is a particularly natural candidate
for the description of strong fluctuations
self-similar over a wide range of scales.
It finds support in the cascade nature, not only of
perturbative QCD, but also of the subsequent hadronization.
However, further work is needed to help understand the details of the
process.
Alternatively, ``classic'' and extensively studied possibilities
are scale-invariant
systems at the critical point of a high-order phase transition.
This subject has attracted particular attention in view of potential
application to quark-gluon plasma formation in heavy-ion collisions.
This paper contains a review of the present status of work on
intermittency and correlations as performed over the last years.
In Chapter~2 we introduce the necessary formalism and collect
useful results and relations widely scattered in the literature.
Chapter~3 describes experimental data
on correlations in various experiments and discusses predictions
of popular models.
Chapter~4 is devoted to
data and models on the subject of particle fluctuations and
the search for power laws.
Chapter~5 gives an overview of the many theoretical ideas
related to the problem of multiplicity scaling and fractality.
Conclusions are summarized in Chapter~6.
\chapter{Formalism}
\section{Definitions and notation}
In this section, we compile and
summarize definitions and various relations
among the physical quantities used in the sequel of this paper.
No originality is claimed in the presentation of this material.
It merely serves the purpose of fixing the notation and assembling
a number of results scattered throughout the literature.
\subsection{Exclusive and inclusive densities}\\
We start by considering a collision between particles a and b yielding
exactly $n$ particles in a sub-volume $\Omega$ of the total phase space
$\Omega_{\mbox{\small tot}}$.
Let the single symbol $y$ represent the kinematical variables needed to
specify the position of each particle in this space (for example, $y$ can be
the c.m. rapidity\footnote{\ The rapidity $y$ is defined as
$y={1\over 2} \ln [(E+p_\ifmath{{\mathrm{L}}})/(E-p_\ifmath{{\mathrm{L}}})]$, with $E$ the energy and $p_\ifmath{{\mathrm{L}}}$ the
longitudinal component of momentum vector $\vec p$ along a given direction
(beam-particles, jet-axis, etc.); pseudo-rapidity is defined as
$\eta={1\over 2}\ln [(p+p_\ifmath{{\mathrm{L}}})/(p-p_\ifmath{{\mathrm{L}}})]$}
variable of each particle and $\Omega$ an interval of length
$\delta y$). The distribution of points in $\Omega$ can be characterized
by continuous probability densities $P_n(y_1,\ldots,y_n)$; $n=1,2,\ldots$.
For simplicity, we assume all final-state particles to be of the same type.
In this case, the {\bf exclusive} distributions $P_n(y_1,\ldots,y_n)$
can be taken fully symmetric in
$y_1,\ldots,y_n$; they describe the distribution in $\Omega$ when the
multiplicity is exactly $n$.
The corresponding {\bf inclusive} distributions are given for
$n=1,2,\ldots$ by:
\begin{eqnarray}
\rho_n(y_1,\ldots,y_n)&=&P_n(y_1,\ldots,y_n)\nonumber\\
&&\mbox{}+\sum_{m=1}^{\infty}\frac{1}{m!} \int_\Omega
P_{n+m}(y_1,\ldots,y_n,{y'}_1,\ldots,{y'}_m)\,\DXX{y'}{m}{i}.
\label{dr:1}
\end{eqnarray}
The inverse formula is
\begin{eqnarray}
P_n(y_1,\ldots,y_n)&=&\rho_n(y_1,\ldots,y_n)\nonumber\\
&&\mbox{}+\sum_{m=1}^{\infty}(-1)^m \frac{1}{m!} \int_\Omega
\rho_{n+m}(\XVEC{y}{n},\XVEC{y'}{m})\,\DXX{y'}{m}{i}\ .
\label{dr:2}
\end{eqnarray}
$\rho_n(y_1,\ldots,y_n)$ is the probability density for $n$ points to be at
$\XVEC{y}{n}$, irrespective of the presence and location of any further
points. The probability $P_0$ of multiplicity zero is given by
\begin{equation}
P_0=1-\sum_{n=1}^{\infty}\frac{1}{n!} \int_\Omega
P_{n}(\XVEC{y}{n})\,\DXX{y}{n}{i}\ .
\label{dr:3}
\end{equation}
This suggests to define $\rho_0=1$ in (\ref{dr:1}).
It is often convenient to summarize the above results with the help of
the generating functional\footnote{\ The technique of generating
functions has been known since Euler's time and was used for functionals by
N.N.~Bogoliubov
in statistical mechanics already in
1946~\cite{bogolubov}; see also~\cite{Brown}}
\begin{equation}
{\cal G}^{{\mathrm{excl}}}\left[z(y)\right]\equiv P_0+\sum_{n=1}^\infty\frac{1}{n!}
\int_\Omega\ P_n(\XVEC{y}{n})\,
z(y_1)\ldots z(y_n)\,\DXX{y}{n}{i}\ ,
\label{dr:4}
\end{equation}
where $z(y)$ is an arbitrary function of $y$ in $\Omega$.
The substitution
\begin{equation}
z(y)=1+u(y)
\label{dr:5}
\end{equation}
gives through (\ref{dr:1}) the alternative expansion
\begin{equation}
{\cal G}^{{\mathrm{incl}}}\left[u(y)\right]=1+\sum_{n=1}^\infty\frac{1}{n!}
\int_\Omega\,\rho_n(\XVEC{y}{n})\,u(y_1)\ldots u(y_n)\DXX{y}{n}{i}
\label{dr:6}
\end{equation}
and the relation
\begin{equation}
{\cal G}^{{\mathrm{incl}}}\left[z(y)\right]={\cal G}^{{\mathrm{excl}}}\left[z(y)+1\right].
\label{dr:7}
\end{equation}
{}From (\ref{dr:4}) and (\ref{dr:7}) one recovers by functional
differentiation:
\begin{equation}
P_n(\XVEC{y}{n})=\left.
\frac{\partial^n {\cal G}^{{\mathrm{excl}}}\left[z(y)\right]}{%
\partial z(y_1)\ldots\partial z(y_n)} \right|_{z=0},
\label{dr:8}
\end{equation}
and
\begin{equation}
\rho_n(\XVEC{y}{n})=\left.
\frac{\partial^n {\cal G}^{{\mathrm{incl}}}\left[u(y)\right]}{%
\partial u(y_1)\ldots\partial u(y_n)} \right|_{u=0}.
\label{dr:9}
\end{equation}
To the set of inclusive number-densities $\rho_n$ corresponds a sequence of
inclusive differential cross sections:
\begin{equation}
\frac{1}{\sigma_{{\mathrm{inel}}}}\,\ifmath{{\mathrm{d}}}\sigma=\rho_1(y)\,\ifmath{{\mathrm{d}}} y,
\label{dr:10}
\end{equation}
\begin{equation}
\frac{1}{\sigma_{{\mathrm{inel}}}} \ifmath{{\mathrm{d}}}^2\sigma=\rho_2(y_1,y_2)\,\ifmath{{\mathrm{d}}} y_1 \ifmath{{\mathrm{d}}} y_2.
\label{dr:10a}
\end{equation}
Integration over an interval $\Omega$ in $y$ yields
\begin{eqnarray}
&~&\int_\Omega \rho_1(y) \ifmath{{\mathrm{d}}} y = \langle n\rangle \nonumber \\
&~&\int_\Omega \int_\Omega \rho_2(y_1,y_2)\ifmath{{\mathrm{d}}} y_1\ifmath{{\mathrm{d}}} y_2 =
\langle n(n-1)\rangle \nonumber \\
&~&\int_\Omega \ifmath{{\mathrm{d}}} y_1 \dots \int_\Omega \ifmath{{\mathrm{d}}} y_q \rho_q (y_1,\dots,y_q) =
\langle n(n-1)\dots (n-q+1)\rangle \ ,
\end{eqnarray}
where the angular brackets imply the average over the event ensemble.
\subsection{Cumulant correlation functions}\\
The inclusive $q$-particle densities $\rho_q(\XVEC{y}{q})$ in general
contain ``trivial'' contributions from lower-order densities. Under
certain conditions, it
is, therefore, advantageous to consider a new sequence of functions
$C_q(\XVEC{y}{q})$ as those statistical quantities which vanish whenever one
of their arguments becomes statistically independent of the others.
It is well known that the quantities with such properties are the
correlation functions--also called (factorial) cumulant functions--or, in
integrated form,
Thiele's semi-invariants~\cite{thiele}. A formal proof of this property was
given by Kubo~\cite{kubo} (see also Chang et al.~\cite{Chang:69}).
The cumulant correlation functions are defined as in the cluster expansion
familiar from statistical mechanics via the sequence
{}~\cite{kahn:uhlenbeck,huang,Mue71}:
\begin{eqnarray}
\rho_1(1)& =& C_1(1),\\
\rho_2(1,2)& =& C_1(1)C_1(2) +C_2(1,2),\\
\rho_3(1,2,3)& =& C_1(1)C_1(2)C_1(3)
+C_1(1)C_2(2,3)
+C_1(2)C_2(1,3)
+\nonumber\\
& &\mbox{}
+C_1(3)C_2(1,2)+C_3(1,2,3);
\end{eqnarray}
and, in general, by
\begin{eqnarray}
\rho_m(1,\ldots,m) &=& \sum_{{\{l_i\}}_m}\sum_{\text{perm.}}
\underbrace{\left[C_1()\cdots C_1()\right]}_{l_1\,\text{factors}}
\underbrace{\left[C_2(,)\cdots C_2(,)\right]}_{l_2\,\text{factors}}
\cdots\nonumber\\
& & \cdots \underbrace{\left[C_m(,\ldots,)\cdots C_m(,\ldots,)
\right]}_{l_m\,\text{factors}}.
\label{a:4}
\end{eqnarray}
Here, $l_i$ is either zero or a positive integer and the sets of integers
$\{l_i\}_m$ satisfy the condition
\begin{equation}
\sum_{i=1}^m i\, l_i=m.\label{a:5}
\end{equation}
The arguments in the $C_i$ functions are to be filled by the $m$ possible
momenta in any order. The sum over permutations is a sum over all
distinct ways of filling these arguments. For any given factor product there
are precisely~\cite{huang}
\begin{equation}
\frac{m!}{
\left[(1!)^{l_1} (2!)^{l_2}\cdots(m!)^{l_m}\right] {l_1!}{l_2!}\cdots{l_m!}}
\label{a:6}
\end{equation}
terms.
The complete set of relations is contained in the functional identity:
\begin{equation}
{\cal G}^{{\mathrm{incl}}}\left[u(y)\right]=\exp{\left\{g\left[u(y)\right]\right\}}\ ,
\label{dr:11}
\end{equation}
where
\begin{equation}
g\left[u(y)\right]=\int\rho_1(y) u(y)\,\ifmath{{\mathrm{d}}} y +\sum_{q=2}^\infty\frac{1}{q!}
\int_\Omega\,C_q(\XVEC{y}{q})\,u(y_1)\ldots u(y_q)\DXX{y}{q}{i}.
\label{dr:12}
\end{equation}
It follows that
\begin{equation}
C_q(\XVEC{y}{q})=
\left.\frac{\partial^q g\left[u(y)\right]}{%
\partial u(y_1)\ldots\partial u(y_n)}\right|_{u=0}.
\label{dr:13}
\end{equation}
The relations (\ref{a:4}) may be inverted with the result:
\begin{eqnarray}
C_2(1,2)&=&\rho_2(1,2) -\rho_1(1)\rho_1(2)\ ,\nonumber\\
C_3(1,2,3)&=&\rho_3(1,2,3)
-\sum_{(3)}\rho_1(1)\rho_2(2,3)+2\rho_1(1)\rho_1(2)\rho_1(3)\ ,\nonumber\\
C_4(1,2,3,4)&=&\rho_4(1,2,3,4)
-\sum_{(4)}\rho_1(1)\rho_3(1,2,3)
-\sum_{(3)}\rho_2(1,2)\rho_2(3,4)\nonumber\\
&&\mbox{} +2\sum_{(6)}\rho_1(1)\rho_1(2)\rho_2(3,4)-6\rho_1(1)
\rho_1(2)\rho_1(3)\rho_1(4).
\label{a:4b}
\end{eqnarray}
In the above relations we have abbreviated $C_q(\XVEC{y}{q})$ to
$C_q(1,2,\ldots,q)$; the summations indicate that all possible permutations
have to be taken (the number under the
summation sign indicates the number of terms).
Expressions for higher orders can be derived from the related formulae given
in~\cite{kendall}.
It is often convenient to divide the functions
$\rho_q$ and $C_q$ by the product of one-particle densities. This leads to
the definition of the normalized inclusive densities and correlations:
\begin{eqnarray}
r_q(\XVEC{y}{q}) &=& \rho_q(\XVEC{y}{q})/\rho_1(y_1)\ldots
\rho_1(y_q),\label{3.8}\\
K_q(y_1,\ldots,y_q)& =& C_q(y_1,\ldots,y_q)/\rho_1(y_1)\ldots
\rho_1(y_q).\label{3.9}
\end{eqnarray}
{}From expression (\ref{dr:11}) it can be deduced that, at finite energy, an
infinite number of $C_q$ will be non-vanishing: The densities
$\rho_q$ vanish for $q>N$, where $N$ is the maximal number of particles
in $\Omega$ allowed e.g. by energy-momentum conservation. As a consequence,
the functional ${\cal G}$ is a ``polynomial'' in $u(y)$. This in turn requires
the exponent in (\ref{dr:11}) to be an ``infinite series'' in $u(y)$.
In other words, the higher-order correlation functions must cancel
the lower-order ones that contribute to a vanishing density function.
Phenomenologically, this implies that it is meaningful to use
correlation functions $C_q$ only if the number of correlated particles
in the considered phase-space domain $\Omega$ is considerably smaller
than the average multiplicity in that region~\cite{Brown}. These
conditions are not always fulfilled in present-day experiments for very
small phase-space cells, with the exception of perhaps $AA$-collisions.
\subsection{Correlations for particles of different species}\\
The generating functional technique of Sect.~2.1.1 can be extended to
the general situation where several different species of particles are
distinguished. This will not be pursued here and we refer to the literature
for details~\cite{Brown,koba,webber:72,eggers:phd}.
Considering two particle species a and b, the two-particle rapidity
correlation function is of the form:
\begin{equation}
C_2^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}(y_1,y_2)= \rho_2^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}(y_1,y_2) -f \rho_1^\ifmath{{\mathrm{a}}}(y_1)
\rho_1^\ifmath{{\mathrm{b}}}(y_2),
\label{dr:ex1}
\end{equation}
with
\begin{equation}
\rho^{\ifmath{{\mathrm{a}}}}_1(y_1)=\frac{1}{\sigma_{{\mathrm{inel}}}}\,\frac{d\sigma^\ifmath{{\mathrm{a}}}}{\ifmath{{\mathrm{d}}} y_1}\,;
\quad
\rho^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}_2(y_1,y_2)=\frac{1}{\sigma_{{\mathrm{inel}}}}\,\frac{\ifmath{{\mathrm{d}}}\sigma^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}}
{\ifmath{{\mathrm{d}}} y_1\ifmath{{\mathrm{d}}} y_2}.
\label{dr:ex2}
\end{equation}
Here, $y_1$ and $y_2$ are the c.m. rapidities, $\sigma_{{\mathrm{inel}}}$ the inelastic
cross section and a, b represent particle properties, e.g. charge.
The normalization conditions are:
\begin{equation}
\int \rho^\ifmath{{\mathrm{a}}}_1(y_1) \ifmath{{\mathrm{d}}} y_1 = \langle n_\ifmath{{\mathrm{a}}} \rangle\;; \int\hskip-2mm\int
\rho^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}_2(y_1, y_2) \ifmath{{\mathrm{d}}} y_1 \ifmath{{\mathrm{d}}} y_2 = \langle n_\ifmath{{\mathrm{a}}}(n_\ifmath{{\mathrm{b}}}-\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}})\rangle\ ,
\label{dr:2.3}
\end{equation}
\begin{equation}
\int\hskip-2mm \int C^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}_2(y_1, y_2) \ifmath{{\mathrm{d}}} y_1 \ifmath{{\mathrm{d}}} y_2 = \langle n_\ifmath{{\mathrm{a}}}
(n_\ifmath{{\mathrm{b}}}-\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}})\rangle - f\langle n_\ifmath{{\mathrm{a}}}\rangle \langle n_\ifmath{{\mathrm{b}}}\rangle \ \ ,
\label{dr:2.4}
\end{equation}
where $\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}=0$ for the case when a and b are particles of different
species and $\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}=1$ for identical particles, and $n_\ifmath{{\mathrm{a}}}$ and $n_\ifmath{{\mathrm{b}}}$
are the corresponding particle multiplicities.
Most experiments use
\begin{equation}
f=1\ \ ,
\label{dr:2.5a}
\end{equation}
so that the integral over the correlation function (equal to the ratio
$\bar n^2/k$ of
the negative binomial parameters~\cite{GiovHove86}) vanishes for the case of
a Poissonian multiplicity distribution.
Other experiments use
\begin{equation}
f = \frac{\langle n_\ifmath{{\mathrm{a}}}(n_\ifmath{{\mathrm{b}}}-\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}})\rangle}{\langle n_\ifmath{{\mathrm{a}}}\rangle\langle n_\ifmath{{\mathrm{b}}}\rangle}
\label{dr:2.5b}
\end{equation}
to obtain a vanishing integral also for a non-Poissonian multiplicity
distribution.
To be able to compare the various experiments, we use both definitions and
denote the correlation function $C_2^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}(y_1,y_2)$ when following
definition (\ref{dr:2.5a}) and $C_2^{'\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}(y_1,y_2)$
when following definition (\ref{dr:2.5b}).
We, furthermore,
use a reduced form of definition (\ref{dr:2.5b}),
\begin{equation}
\tilde C^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}_2(y_1,y_2)=C'^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}_2(y_1,y_2)/\langle n_\ifmath{{\mathrm{a}}}(n_\ifmath{{\mathrm{b}}}-
\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}})\rangle.
\label{dr:2.6}
\end{equation}
\noindent
The corresponding normalized correlation functions
\begin{equation}
K^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}_2 (y_1,y_2) = {C^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}}_2(y_1,y_2)\over f \rho^\ifmath{{\mathrm{a}}}_1(y_1)
\rho^\ifmath{{\mathrm{b}}}_1(y_2)}
\label{dr:2.7}
\end{equation}
follow the relations
\begin{equation}
{K_2}' = {1\over f} (K_2+1) - 1\ \ ,
\label{dr:2.8}
\end{equation}
and $\tilde K_2$ is defined as $\tilde K_2={K_2}'$.
These are more appropriate than $C_2$ when comparisons have to be performed
at different average multiplicity and are less sensitive to acceptance
problems.
The correlation functions defined by expressions
(\ref{dr:ex1})-(\ref{dr:2.8}), contain a
pseudo-correl\-ation due to the summation of events with different charge
multiplicity $n$ and different
semi-inclusive single-particle densities $\rho^{(n)}_1$.
The relation between inclusive and semi-inclusive correlation functions has
been carefully analyzed in~\cite{Carr90}. Let $\sigma_n$ be the topological
cross section and
\begin{equation}
P_n = \sigma_n/\Sigma\sigma_n\ \ .
\end{equation}
The semi-inclusive rapidity single- and two-particle densities
for particles a and b are defined as
\begin{equation}
\rho^{(n)}_1(y)={1\over \sigma_n}{\ifmath{{\mathrm{d}}}\sigma^a_n\over \ifmath{{\mathrm{d}}} y}\ \ {\rm and}\ \
\rho^{(n)}_2(y_1,y_2)={1\over \sigma_n} {\ifmath{{\mathrm{d}}}\sigma^{ab}_n\over \ifmath{{\mathrm{d}}} y_1\ifmath{{\mathrm{d}}} y_2}\ .
\label{dr:2.12}
\end{equation}
The inclusive correlation function $C_2(y_1,y_2)$ can then be written as
\begin{equation}
C_2(y_1,y_2) = C_\ifmath{{\mathrm{S}}} (y_1,y_2) + C_\ifmath{{\mathrm{L}}}(y_1,y_2)\ ,
\label{dr2.9}
\end{equation}
where
\begin{equation}
C_\ifmath{{\mathrm{S}}}(y_1,y_2) = \Sigma P_n C_2^{(n)} (y_1,y_2) \label{cs37}
\end{equation}
\begin{equation}
C_\ifmath{{\mathrm{L}}}(y_1,y_2) = \Sigma P_n \Delta\rho^{(n)}(y_1)\Delta\rho^{(n)}(y_2)
\end{equation}
with $C^{(n)}_2(y_1,y_2) = \rho_2^{(n)}(y_1,y_2) - \rho_1^{(n)}(y_1)
\rho_1^{(n)}(y_2)$ and $\Delta\rho^{(n)}(y)=\rho_1^{(n)}(y)-\rho_1(y)$. In
{}~(\ref{cs37}) $C_\ifmath{{\mathrm{S}}}$ is the average of the semi-inclusive correlation
functions (often misleadingly denoted as ``short-range") and is more
sensitive to dynamical correlations. The term $C_\ifmath{{\mathrm{L}}}$ (misleadingly called
``long-range") arises from mixing different topological single-particle
densities.
A normalized form of $C_\ifmath{{\mathrm{S}}}$ can be defined as
\begin{equation}
K_{\ifmath{{\mathrm{S}}}} (y_1,y_2) = {C_{\ifmath{{\mathrm{S}}}}(y_1,y_2)\over \sum_nP_n\rho^{(n)}_1(y_1)
\rho^{(n)}_1(y_2)}={\sum_nP_n\rho^{(n)}_2(y_1,y_2)\over
\sum_nP_n\rho^{(n)}_1(y_1)\rho^{(n)}_1(y_2)} - 1
\ \ .\label{dr:2.13}
\end{equation}
$C'_{\ifmath{{\mathrm{S}}}}$ and $\tilde C_{\ifmath{{\mathrm{S}}}}$ and their normalized forms $K'_{\ifmath{{\mathrm{S}}}}$ and
$\tilde K_{\ifmath{{\mathrm{S}}}}$ are defined accordingly, with the averages $\langle n\rangle$
and $\langle n_\ifmath{{\mathrm{a}}}(n_\ifmath{{\mathrm{b}}}-\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}})\rangle$ replaced by $n$ and
$n_\ifmath{{\mathrm{a}}}(n_\ifmath{{\mathrm{b}}}-\delta^{\ifmath{{\mathrm{a}}}\ifmath{{\mathrm{b}}}})$,
respectively.
Analogous expressions may be derived for three-particle correlations. They
are discussed in Sect.~3.4.
\subsection{Factorial and cumulant moments}\\
When the parametric function $z(y)$ is replaced by a constant $z$, the
generating functionals reduce to the generating function for
the multiplicity distribution.
Indeed, the probability $P_n$ for producing $n$ particles is
given by
\begin{equation}
P_n=\sigma_n^{{\mathrm{excl}}}/\sigma_{{\mathrm{inel}}}
\label{dr:14}
\end{equation}
and we have
\begin{eqnarray}
G(z)&=&\sum_{n=0}^\infty P_n(1+z)^n={\cal G}^{{\mathrm{excl}}}[z+1]=
{\cal G}^{{\mathrm{incl}}}[z]\\
&=& 1 +\sum_{q=1}^\infty \frac{z^q}{q!}\int_\Omega\rho_q(\XVEC{y}{q})\,
\ifmath{{\mathrm{d}}} y_1\ldots \ifmath{{\mathrm{d}}} y_q\\
&=& 1 +\sum_{q=1}^\infty \frac{z^q}{q!}\;\tilde{F}_q\ .
\label{dr:15}
\end{eqnarray}
The $\tilde{F}_q$ are the unnormalized factorial (or binomial) moments
\begin{eqnarray}
\tilde F_q \ \equiv \ \FMM{n}{q}&\equiv& \aver{n(n-1)\ldots(n-q+1)}\nonumber\\
&=& \int_\Omega \ifmath{{\mathrm{d}}} y_1\ldots\int_\Omega \ifmath{{\mathrm{d}}} y_q\;
\rho_q(\XVEC{y}{q})\nonumber\\
&=& \sum_n P_n\, n(n-1)\ldots(n-q+1).
\label{dr:16}
\end{eqnarray}
This relation can (formally) be inverted. If $P_n=0$ for
$n>N$ then an approximation for $P_n$ is given by:
\begin{equation}
P_n=\frac{1}{n!}\sum_{j=0}^{N-n} (-1)^j \frac{\tilde{F}_{j+n}}{j!}
\quad (n=0,1,\ldots N),
\label{dr:28b}
\end{equation}
and $P_n$ is included between any two successive values obtained by
terminating the sum at $j=s$ and $j=s+1$, respectively.
In (\ref{dr:16}) $n$ denotes the multiplicity in $\Omega$ and the average is
taken over
the ensemble of events. All the integrals are taken over the
same volume $\Omega$ such that $y_i\in\Omega$ $\forall i\in\{1,\ldots,q\}$.
Using the correlation-function cluster decomposition, one further has
\begin{equation}
\log G(z)=\aver{n}z +\sum_{q=2}^{\infty} \frac{z^q}{q!}\; f_q.
\label{dr:17}
\end{equation}
The $f_q$ are the unnormalized factorial cumulants, also known as Mueller
moments~\cite{Mue71}
\begin{equation}
f_q=
\int_\Omega \ifmath{{\mathrm{d}}} y_1\ldots\int_\Omega \ifmath{{\mathrm{d}}} y_q\, C_q(\XVEC{y}{q}),
\label{dr:18}
\end{equation}
the integrations being performed as in (\ref{dr:16}).
The quantities $\tilde{F}_q$ and $f_q$ are easily found if $G(z)$ is
known:
\begin{eqnarray}
\tilde{F}_q&=& \left.\frac{\ifmath{{\mathrm{d}}}^q G(z)}{\ifmath{{\mathrm{d}}} z^q}\right|_{z=0}\ \ \ ,\\
f_q &=& \left.\frac{\ifmath{{\mathrm{d}}}^q \log{G(z)}}{\ifmath{{\mathrm{d}}} z^q}\right|_{z=0}\\
\noalign{and}
P_q&=& \frac{1}{q!}\left.\frac{\ifmath{{\mathrm{d}}}^q G(z)}{\ifmath{{\mathrm{d}}} z^q}\right|_{z=-1}.
\end{eqnarray}
Using Cauchy's theorem, this can also be written as
\begin{equation}
P_n=\frac{1}{2\pi i}\oint\frac{G(z)}{(1+z)^{n+1}}\,\ifmath{{\mathrm{d}}} z,
\label{dr:33a}
\end{equation}
where the integral is on a circle enclosing $z=-1$.
Equation~(\ref{dr:33a}) is sometimes useful in deriving asymptotic
expressions for $P_n$ in terms of factorial moments or
cumulants~\cite{weisberger,Mue71}.
As a simple example, we consider the Poisson distribution
$$P_n = \ifmath{{\mathrm{e}}}^{-\aver{n}}\frac{\aver{n}^n}{n !}\ \ \ ,$$
for which
\begin{equation}
G(z)=\sum_0^\infty P_n\,(1+z)^n=\exp{\{\aver{n}z\}}\ \ \ ,
\label{dr:19}
\end{equation}
showing that $f_q\equiv0$ for $q>1$.
In that case one has:
\begin{equation}
\tilde{F}_q=\aver{n(n-1)\ldots(n-q+1)}=\aver{n}^q.
\label{dr:ex3}
\end{equation}
The expressions of density functions in terms of cumulant correlation
functions, and the reverse relations,
are duplicated for their integrated counterparts. They
follow directly from the equations:
\begin{eqnarray}
1+\sum_{q=1}^{\infty} \frac{z^q}{q!}\,\tilde{F}_q&=&\exp{\{\aver{n}z +
\sum_{q=2}^\infty \frac{z^q}{q!}f_q \}}\label{dr:37}\\ \noalign{or\hfill}
\log\left(1+\sum_{q=1}^{\infty} \frac{z^q}{q!}\,\tilde{F}_q\right)&=&
\aver{n}z+\sum_{q=2}^{\infty} \frac{z^q}{q!} f_q\label{dr:38}
\end{eqnarray}
by expanding either the exponential in (\ref{dr:37}) or the logarithm
in (\ref{dr:38}) and equating the
coefficients of the same power of $z$. One finds~\cite{kendall}:
\begin{eqnarray}
\tilde{F}_1&=& f_1\ \ \ , \nonumber\\
\tilde{F}_2&=& f_2+f_1^2\ \ \ ,\nonumber\\
\tilde{F}_3&=& f_3+ 3f_2f_1+f_1^3\ \ \ ,\nonumber\\
\tilde{F}_4&=& f_4 + 4f_3f_1 + 3f_2^2 +6f_2f_1^2 +f_1^4\ \ \ ,\nonumber\\
\tilde{F}_5&=& f_5 + 5f_4f_1 +10 f_3f_2 +10f_3f_1^2 +15f_2^2 f_1 +10f_2f_1^3
+f_1^5\ ;
\label{dr:37a}
\end{eqnarray}
and in general:
\begin{equation}
{\tilde{F}_q} = q! \sum_{{\{l_i\}}_q} \prod_{j=1}^q
\left(\frac{{f_j} }{j!}\right)^{l_j} \frac{1}{l_j!}\ , \label{dr:39}
\end{equation}
with the summation as in (\ref{a:4}) and
$\sum_{l=1}^q i\, l_i=q$.
The latter formula can also be written as:
\begin{equation}
\tilde{F}_q=\sum_{l=0}^{q-1} \left(\!\begin{array}{c}
q-1\\ l \end{array} \!\right)\,f_{q-l}\,\tilde{F}_l \ ,
\label{dr:20}
\end{equation}
(with $\tilde{F}_0\equiv1$, $f_0\equiv0$)
and is well-suited for computer calculation.
An equivalent relation was derived in~\cite{Mue71}.
The (ordinary) moments:
\begin{equation}
\mu_q=\aver{n^q}=\sum_{n=0}^{\infty} n^q\,P_n\ \ ,
\label{dr:39a}
\end{equation}
may be derived from the moment generating function
\begin{equation}
M(z)=\sum_{n=0}^\infty \ifmath{{\mathrm{e}}}^{\textstyle nz}P_n\ \ ,
\label{dr:40}
\end{equation}
since
\begin{equation}
\mu_q=\left.\frac{\ifmath{{\mathrm{d}}}^qM(z)}{\ifmath{{\mathrm{d}}} z^q}\right|_{z=0}.
\label{dr:41}
\end{equation}
We note the useful relations
\begin{eqnarray}
M(z)&=&G\left(\ifmath{{\mathrm{e}}}^{\textstyle z}-1\right)\ \ ,\\
G(z)&=&M\left(\log(1+z)\right)\ \ .
\label{dr:41a}
\end{eqnarray}
Moments and factorial moments are related to each other by series
expansions.
{}From the identities~\cite{abramowitz}:
\begin{eqnarray}
n(n-1)\ldots(n-q+1)&=&\sum_{m=0}^{q}S_q^{(m)}\,n^m\ \ ,\\
n^q&=&\sum_{m=0}^q{\cal S}_q^{(m)}\,n(n-1)\ldots(n-m+1)\ \ ,
\label{dr:42}
\end{eqnarray}
where $S_q^{(m)}$ and ${\cal S}_q^{(m)}$ are Stirling numbers of the first
and second kind, respectively, follows directly:
\begin{eqnarray}
\tilde{F_q}&=&\sum_{m=0}^{q}S_q^{(m)}\;\mu_m\ \ ,\\
\mu_q&=&\sum_{m=0}^q{\cal S}_q^{(m)}\;\tilde{F}_m\ \ .
\label{dr:43}
\end{eqnarray}
Cumulants $\kappa_q$ can be defined in terms of the moments $\mu_q$ in
the standard way~\cite{cramer,kendall}. They obey relations identical to
(\ref{dr:37a}). The cumulants are integrals of the type (\ref{dr:18}) of
differential quantities known as density moments. These are discussed in
{}~\cite{weingarten,koba:weingarten}. Relations expressing central moments
in terms of factorial moments via non-central Stirling numbers are
derived in~\cite{Kout82}.
\subsection{Cell-averaged factorial moments and cumulants;\\
generalized moments}\\
In practical work, with limited statistics,
it is almost always necessary to perform
averages over more than a single phase-space cell. Let $\Omega_m$
be such a cell (e.g. a single rapidity interval of size $\delta y$) and
divide the phase-space volume into $M$ non-overlapping cells $\Omega_m$ of
size $\delta\Omega$, independent of $m$.
Let $n_m$ be the number of particles in cell $\Omega_m$.
Different cell-averaged moments may be considered, depending
on the type of averaging.
Normalized factorial moments~\cite{bialas1,bialas2},
which have become known as {\it vertical moments},
are defined as\footnote{\ Here and in the following we consider rapidity space
for definiteness}
\begin{eqnarray}
F^\ifmath{{\mathrm{V}}}_q(\delta y)&\equiv&
\frac{1}{M}\;\sum_{m=1}^M \frac{\aver{\FACT{n}{q}}}{\aver{n_m}^q}
\label{dr:44}\\
&\equiv & \frac{1}{M} \sum^M_{m=1}
\frac{\int_{\delta y} \rho_q(y_1,\ldots,y_q) \prod^q_{i=1} \ifmath{{\mathrm{d}}} y_i}
{\left(\int_{\delta y} \rho(y)dy\right)^q}\ \ , \nonumber \\
&= & \frac{1}{M(\delta y)^q} \sum^M_{m=1} \int_{\delta y} \frac{\rho_q(y_1,\dots,y_q)
\prod^q_{i=1} \ifmath{{\mathrm{d}}} y_i}{\left(\bar\rho_m\right)^q}. \label{dr:45}
\end{eqnarray}
The full rapidity interval $\Delta Y$ is divided into $M$ equal bins:
$\Delta Y=M\delta y$; each $y_i$ is within the $\delta y$-range and
$\aver{n_m}\equiv\overline{\rho}_m\delta y \equiv \int_{\delta y} \rho_1(q)\ifmath{{\mathrm{d}}} y$.
One may also define normalized {\it horizontal moments}
by
\begin{equation}
F^\ifmath{{\mathrm{H}}}_q(\delta y)\equiv \frac{1}{M}\;\sum_{m=1}^M
\frac{\aver{\FACT{n}{q}}}{\aver{\overline{n}_m}^q}\ \ .
\label{dr:46}
\end{equation}
with $\overline{n}_m=\sum_m n_m/M;\quad \aver{\overline{n}_m}=\aver{n}/M;
\quad n=\sum_mn_m$.
Horizontal and vertical moments are equal if $M=1$. Vertical moments
are normalized locally and thus sensitive only to fluctuations within each
cell but not to the overall shape of the single-particle density.
Horizontal moments are sensitive to the shape of the single-particle density
in $y$ and further depend on the correlations between cells.
To eliminate the effect of a non-flat rapidity distribution,
it was suggested to either introduce correction factors~\cite{Fial89}
or use ``cumulative'' variables which transform
an arbitrary distribution into a uniform one~\cite{Ochs,BiGa90}.
Likewise, cell-averaged normalized factorial cumulant moments may be
defined as
\begin{equation}
K_q(\delta y) = \frac{1}{M(\delta y)^q}
\sum^M_{m=1} \int\limits_{\delta y} \prod_i\ifmath{{\mathrm{d}}} y_i
\frac{C_q(y_1,\ldots,y_q)}{(\bar\rho_{m})^q}\ \ \ . \label{dr:47}
\end{equation}
They are related~\cite{CaES91} to the factorial moments by\footnote{\ The
higher-order relations can be found in~\cite{CaES91}}
\begin{eqnarray}
F_2&=&1+K_2\ \ \ ,\nonumber\\
F_3&=&1+3K_2+K_3\ \ \ ,\nonumber\\
F_4&=&1+6K_2+3\overline{K^2_2}+4K_3+K_4 \,.\label{dr:48}
\end{eqnarray}
In $F_4$ and higher-order moments, ``bar averages" appear. They are defined
as $\overline{AB}\equiv\sum\limits_m A_mB_m/M$.
Besides factorial and cumulant moments, other measures of multiplicity
fluctuations have been proposed. In particular,
$G$-moments~\cite{Feder88}---known in statistics as
frequency moments~\cite{kendall}---were
extensively used to investigate
whether multiparticle processes possess
(multi)fractal properties~\cite{Hwa90,Hwa90-91}.
$G$-moments are defined as
\begin{equation}
G_q=\sum^{M}_{m=1}\hskip-1mm ' p^q_m , \hs2truecm p_m=n_m/n, \hs2truecm
n=\sum^M_{m=1}n_m\ \ .
\label{gghwa}
\end{equation}
Also here, $n_m$ is the number of particles in bin $m$,
the absolute frequency; $n$ is the total
multiplicity in an initial interval
and $M$ is the number of bins at ``resolution'' $M$.
Bins with zero content (``empty bins'')
are excluded in the sum, so that $q$ can cover the whole
spectrum of real numbers. For $q$ negative, $G_q$ is sensitive to ``holes"
in the rapidity distribution of a single event.
Note that $p_m$ in (\ref{gghwa}) is not a probability but
a relative frequency or ``empirical measure'' in
modern terminology. For small $n$, $G$-moments are very sensitive to
statistical fluctuations (``noise''), especially for large M.
This seriously limits their potential.
In attempts to reduce this noise-sensitivity, modified definitions have
been proposed in~\cite{Flor91}.
\subsection{Multivariate distributions}\\
The univariate factorial moments $\tilde{F}_q$
characterize multiplicity fluctuations in a single phase-space cell and thus
reflect only local properties. More information is contained in the
correlations between fluctuations (within the same event) in two or more
cells. This has led to consider multivariate factorial moments. For
non-overlapping cells, the 2-fold factorial moments, also called
{\it correlators}, are defined as:
\begin{equation}
\tilde{F}_{pq}=\aver{n^{[p]}_m {n}^{[q]}_{m'}},
\label{f4:1}
\end{equation}
where $n_m$ ($n_{m'}$) is the number of particles in cell $m$ (cell $m'$). A
normalized version of the two-fold {correlator} is discussed in
{}~\cite{bialas1} and defined as:
\begin{equation}
F_{pq}=\frac{\langle n^{[p]}_m n^{[q]}_{m'}\rangle}{\tilde{F}_p \tilde{F}_q}\ .
\label{f4:2}
\end{equation}
For reasons of statistics, these quantities are usually averaged over many
pairs of cells, keeping the ``distance'' ($D$) between the cells
constant\footnote{\ In one-dimensional rapidity space, $D$ is
defined as the distance between the centers of two rapidity intervals; in
multidimensional phase space a proper metric must first be defined}.
This averaging procedure requires the same precautions regarding
stationarity of single particle densities as for their single-cell
equivalents.
Multi-fold factorial moments are a familiar tool in radio- and
radar physics and in quantum optics~\cite{saleh}. There, they relate to
simultaneous measurement of photo-electron counts
detected in, say $M$, time-intervals,
or in $M$ space points, leading to a joint probability distribution
$P_M(n_1,n_2,\ldots,n_M)$. The importance of
multi-fold moments derives from the fact that, e.g. in the
simplest case of two cells, $\tilde{F}_{11}=\aver{n_m n_{m'}}$ is directly
related to the auto-correlation function of the radiation field and obeys,
for small cells, the Siegert-relation~\cite{saleh}, whatever the
statistical properties of the field.
The higher-order moments are sensitive to higher-order correlations
and to the phase of the field.
Factorial moments and factorial correlators are intimately related
quantities. In terms of inclusive densities one has:
\begin{equation}
\tilde{F}_{pq}=
\int_{\Omega_1}\,\ifmath{{\mathrm{d}}} y_1\ldots \ifmath{{\mathrm{d}}} y_p \int_{\Omega_2} \ifmath{{\mathrm{d}}} y_{p+1}\ldots
\ifmath{{\mathrm{d}}} y_{p+q}\, \rho_{p+q}(y_1,\ldots,y_p;y_{p+1},\ldots,y_{p+q}),\label{f4:3}
\end{equation}
where $\rho_{p+q}$ is the inclusive density of order $p+q$.
The integrations are performed over two arbitrary (possibly overlapping)
phase-space cells
$\Omega_1$ and $\Omega_2$, separated by a ``distance'' $D$.
It should be noted that
the definition (\ref{f4:3}) is more general than (\ref{f4:1}).
For $\Omega_1=\Omega_2$ or $D=0$, (\ref{f4:3}) reduces to the
correct definition of $\tilde{F}_2$
whereas (\ref{f4:1}) is, in this case,
equal to $\aver{n^2}$ and misses the so-called
``shot-noise'' term $-\aver{n}$.
Factorial moments and factorial correlators of the same order
are thus seen to differ only in the choice of the integration domains.
Note that for $p\neq q$, definition (\ref{f4:3}) is not symmetric
in $p$ and $q$ and a symmetrized version is often used in experimental work:
\begin{equation}
\tilde{F}^{(s)}_{pq}=(\tilde{F}_{pq}+\tilde{F}_{qp})/2\,.
\label{dr:49}
\end{equation}
{}From (\ref{f4:3}) follows that $F_{11}$ is directly derivable from measured
two-particle correlation functions or from appropriate analytical
parametrizations. Higher-order correlators involve higher-order
density-functions which, in general, are unknown.
We now turn to a discussion of multivariate factorial cumulants.
For $M$ non-overlapping cells, we introduce the $M$-variate multiplicity
distribution $P_M(\XVEC{n}{M})$ and the corresponding moment- and
factorial-moment generating functions:
\begin{eqnarray}
\hskip-7mm M(\XVEC{z}{M})&=&\sum_{n_1=0}^\infty\sum_{n_2=0}^\infty\cdots
\sum_{n_M=0}^\infty \;e^{z_1n_1+\cdots+ z_Mn_M}\,P_M(\XVEC{n}{M})\ \ ,
\label{dr:50}\\
\hskip-7mm G(\XVEC{z}{M})&=&\sum_{n_1=0}^\infty\sum_{n_2=0}^\infty\cdots
\sum_{n_M=0}^\infty
\;(1+z_1)^{n_1}\cdots(1+z_M)^{n_M}\,P_M(\XVEC{n}{M})\ \ ,\label{dr:51}
\end{eqnarray}
from which the $M$-variate moments are easily obtained by differentiation:
\begin{eqnarray}
\hskip-7mm \mu_{q_1\dots q_M}=\aver{n_1^{q_1}\dots n_M^{q_M}}&=&
\left.\left(
\frac{\partial}{\partial z_1}\right)^{q_1}\cdots\left(
\frac{\partial}{\partial z_M}\right)^{q_M}\; M(\XVEC{z}{M})
\right|_{z_1=\cdots z_M=0} , \label{dr:52}\\
\hskip-7mm \tilde{F}_{q_1\dots q_M}=\aver{n_1^{[q_1]}\dots n_M^{[q_M]}}&=&
\left. \left( \frac{\partial}{\partial z_1}\right)^{q_1}\cdots\left(
\frac{\partial}{\partial z_M}\right)^{q_M}\; G(\XVEC{z}{M})
\right|_{z_1=\cdots z_M=0} . \label{dr:53}
\end{eqnarray}
The multivariate (ordinary) cumulants $\kappa_{q_1\dots q_M}$ and multivariate
factorial cumulants $f_{q_1\dots q_M}$ are likewise
obtained by replacing $M(.)$ and $G(.)$ in (\ref{dr:52}) and
(\ref{dr:53}) by their respective natural logarithms~\cite{Cantrell:70}.
The same expressions serve to extend the relations between univariate
moments and cumulants to their multivariate counterparts.
For $M=2$ and non-overlapping cells, one has the identity
[cfr.~(\ref{dr:37})]:
\begin{equation}
\sum_{l=0}^{\infty}\sum_{m=0}^{\infty} \frac{(z_1)^l(z_2)^m}{l!\,m!}
\tilde{F}_{lm}=
\exp{( \sum_{l=0}^{\infty} \sum_{m=0}^{\infty}
\frac{(z_1)^l(z_2)^m}{l!\,m!} f_{lm})},\label{s1:14}
\end{equation}
where $\tilde{F}_{00}\equiv1$ and $f_{00}$ is defined equal to zero.
It follows that\footnote{\ See also~\cite{eggers:correlators}}
\begin{eqnarray}
\tilde{F}_{11} & = & f_{11}+f_{01} f_{10}, \label{eq:b7} \\
\tilde{F}_{12} & = &
f_{12}
+f_{01} f_{20}
+2 f_{10} f_{11}
+f_{01} f_{10}^2, \label{eq:b8} \\
\tilde{F}_{13} & = &
f_{13}
+f_{01} f_{30}
+3 f_{11} f_{20}
+3 f_{01} f_{10} f_{20}
+3 f_{10} f_{12}
+3 f_{10}^2 f_{11}
+f_{01} f_{10}^3, \label{eq:b9} \\
\tilde{F}_{2 2} & = &
f_{2 2}
+2 f_{10} f_{2 1}
+f_{02} f_{20}
+f_{01}^2 f_{20}
+2 f_{01} f_{12}
+2 f_{11}^2
+4 f_{01} f_{10} f_{11}
\nonumber\\& &\mbox{}%
+f_{02} f_{10}^2
+f_{01}^2 f_{10}^2\ \ . \label{eq:b10}
\end{eqnarray}
Similarly, expanding the logarithm in
\begin{equation}
\sum_{l=0}^{\infty}\sum_{m=0}^{\infty} \frac{(-z_1)^l(-z_2)^m}{l!\,m!}
f_{lm}=
\log{\left(\sum_{l=0}^{\infty} \sum_{m=0}^{\infty}
\frac{(-z_1)^l(-z_2)^m}{l!\,m!} \tilde{F}_{lm}\right)},\label{s1:14bis}
\end{equation}
in powers of $s$ and $t$
and identifying coefficients, the reverse relations follow:
\begin{eqnarray}
\hskip-7mm f_{11}& = &
\tilde{F}_{11}-\tilde{F}_{01}\tilde{F}_{10}\ \ ,\\
\hskip-7mm f_{12}& = &
\tilde{F}_{12}
-\tilde{F}_{01}\tilde{F}_{20}
-2\tilde{F}_{10}\tilde{F}_{11}
+2\tilde{F}_{01}\tilde{F}_{10}^2 \ \ ,\\
\hskip-5mm f_{13}& = &
\tilde{F}_{13}
-\tilde{F}_{01}\tilde{F}_{30}
-3\tilde{F}_{11}\tilde{F}_{20}
+6\tilde{F}_{01}\tilde{F}_{10}\tilde{F}_{20}
-3\tilde{F}_{10}\tilde{F}_{12}
+6\tilde{F}_{10}^2\tilde{F}_{11}
-6\tilde{F}_{01}\tilde{F}_{10}^3 \ \ ,\\
\hskip -7mm f_{22}& = &
\tilde{F}_{22}
-2\tilde{F}_{10}\tilde{F}_{21}
-\tilde{F}_{02}\tilde{F}_{20}
+2\tilde{F}_{01}^2\tilde{F}_{20}
-2\tilde{F}_{01}\tilde{F}_{12}
-2\tilde{F}_{11}^2
\nonumber\\& &\mbox{}%
+8\tilde{F}_{01}\tilde{F}_{10}\tilde{F}_{11}
+2\tilde{F}_{02}\tilde{F}_{10}^2
-6\tilde{F}_{01}^2\tilde{F}_{10}^2\ \ .
\end{eqnarray}
The quantities $\tilde{F}_{0i}$, $\tilde{F}_{i0}$,
$f_{0i}$ and $f_{i0}$ are equal to the single-cell
factorial moments and factorial cumulants, respectively.
Expressions for $\tilde{F}_{ji}$ ($f_{ji}$) are obtained from
the corresponding expression for $\tilde{F}_{ij}$
($f_{ij}$) by permutation of the subscripts. By definition,
$f_{01}=\tilde{F}_{01}$ and $f_{01}$ is equal to the average multiplicity
in cell 2.
It may be noted that the bivariate relations reduce to the
univariate ones (\ref{dr:37a}) by simply amalgamating the indices. For
example, from
\begin{equation}
\tilde{F}_{12} =
f_{12}
+f_{01} f_{20}
+2 f_{10} f_{11}
+f_{01} f_{10}^2, \label{eq:b8a}
\end{equation}
one recovers, by summing the indices
\begin{equation}
\tilde{F}_3=f_3 +3 f_1f_2 +f_1^3.
\label{dr:54}
\end{equation}
It is shown in~\cite{kendall} (Sect.~13.12) that the above relations, while
seemingly complex, have in fact a surprisingly elegant structure, rooted in
simple algebraic properties of completely symmetric functions. Further
discussion on this point and other useful properties may be found
in~\cite{Cantrell:70}.
Extensions to more than two cells is straightforward, in principle, but
involves tedious algebra.
\section{Poisson-noise suppression}
To detect dynamical fluctuations in the density of particles produced in a
high-energy collision, a way has to be devised to eliminate, or to reduce as
much as possible, the statistical fluctuations---noise---due to the finiteness
of the number of particles in the counting cell(s). This requirement can to a
large extent be satisfied by studying
factorial moments and their multivariate
counterparts. It forms the basis of the factorial moment technique, known in
optics, but rediscovered in multi-hadron physics in~\cite{bialas1,bialas2}.
The method rests on the conjecture that the multi-cell
multiplicity distribution $P_M(\XVEC{n}{M})$ can be written as
\newcommand{\PMN}{%
\int \ifmath{{\mathrm{d}}}\rho_1\ldots \ifmath{{\mathrm{d}}}\rho_M P_{\rho}(\rho_1,\ldots,\rho_M)
\prod_{m=1}^M\frac{(\rho_m \delta)^{n_m}}{n_m!}\exp{(-\rho_m \delta)}
}
\begin{equation}
P_M(n_1,\cdots,n_M)= \PMN . \label{s1:1}
\end{equation}
The Poisson factors represent uncorrelated fluctuations of $n_m$
around the average $\rho_m \delta=\aver{n_m}$ in $m$-th interval; $\delta$ is here
the size of the interval.
This can also be written as:
\begin{equation}
P_M(n_1\cdots n_M)=
\aver{\prod_{m=1}^M \frac{\aver{n_m}^{n_m}}{n_m!}
\exp{(-\aver{n_m})}}_\rho\ , \label{s1:2}
\end{equation}
where the outer brackets mean that an average is taken over the probability
distribution of the
densities $\rho_m$, which are subject only to dynamical fluctuations.
If these are absent, $P_{\rho}(\rho_1,\ldots,\rho_M)$ is simply a
product of $\delta$-functions.
The formulae (\ref{s1:1}-\ref{s1:2}) are formally identical to
the expression for the multi-interval photo-electron counting
probability distribution in quantum optics and based
on the famous Mandel formula~\cite{Mandel:58:1,Mandel:59:1}. The latter
relates the probability distribution of {\em the number of detected
photo-electrons} to the statistical distribution of the {e.m. field}.
In optics, $\rho_m$ has the meaning of a space- or time-integrated
field intensity.
The ensemble
average is calculated from the field density matrix which
describes its statistical properties.
Equations (\ref{s1:1})-(\ref{s1:2}) express $P_M(n_1,\ldots,n_M)$ as
a linear transformation of \break
$P_{\rho}(\rho_1,\ldots,\rho_M)$ with a ``Poisson kernel''.
This transformation is known as the\break
``Poisson Transform'' of $P_{\rho}$~\cite{Wolf:64}.
The Poisson-transform of a single-variable function $f(x)$ is the function
$\tilde{f}(n)$ ($n$ integer) defined by the linear transformation
\begin{equation}
\tilde{f}(n) =\int_0^\infty dx\,f(x) \frac{x^n}{n!} e^{-x}.\label{a:pt1}
\end{equation}
A trivial example is the function $\delta(x-\mu)$ whose transform
is the Poisson probability distribution.
The Bose-Einstein distribution
\begin{equation}
\tilde{f}(n)= \frac{\mu^n}{(1+\mu)^{n+1}} \ \ \ \ (n=0,1,\ldots),\label{a:pt2}
\end{equation}
is obtained as the Poisson-transform of the exponential function
$(1/\mu) \exp(-x/\mu)$.
For suitably-behaved functions, the inverse Poisson-transform
exists. It is closely related to the Laplace-transform of $f(x)$.
Several practical methods have been developed to determine the
function $f(x)$ from its Poisson-transform. Besides
methods based on series expansions, the inversion
problem may be reduced to an inverse moment problem.
This follows from the equality between the factorial moments of
$\tilde{f}(n)$ and the ordinary moments of $f(x)$, as further discussed
below.
A table of useful transforms for probability distributions
and further mathematical properties can be found in~\cite{saleh}.
{}From the basic Poisson transform equation (\ref{s1:2}) it is easily seen
that
the multi-fold factorial moment generating function has the simple form
\begin{equation}
G(\XVEC{z}{M}) = \aver{\prod_{j=1}^M
\exp{(z_j\rho_j \delta)}}_\rho,\label{s1:7}
\end{equation}
where the statistical average is again taken over the ensemble of densities
$\rho_1,\cdots,\rho_M$, as indicated by the subscript.
On the other hand, the (ordinary) moment generating function of the
densities is given by:
\begin{eqnarray}
Q(\XVEC{z}{M})&=&\int P_\rho(\XVEC{\rho}{M})
\ifmath{{\mathrm{e}}}^{\rho_1z_1+\cdots+\rho_Mz_M}\ifmath{{\mathrm{d}}}\rho_1\ldots \ifmath{{\mathrm{d}}}\rho_M\\
&=&\left\langle\prod_{j=1}^M \ifmath{{\mathrm{e}}}^{\rho_jz_j}\right\rangle_{\textstyle
\rho}\ \ .
\label{dr:55}
\end{eqnarray}
Comparing (\ref{s1:7}) and (\ref{dr:55}), it follows that:
\begin{equation}
G(\XVEC{z}{M})=Q(\delta\rho_1z_1,\ldots,\rho_Mz_M\delta)\ \ .
\label{dr:56}
\end{equation}
This equation implies that the normalized multi-variate factorial moments
of the multiplicity distribution
\begin{equation}
F_{q_1\ldots q_M}=\frac{%
\tilde{F}_{q_1\ldots q_M}}{%
\aver{n_1}^{q_1}\ldots\aver{n_M}^{q_M}}\
\label{dr:56a}
\end{equation}
are equal to the
normalized multivariate (ordinary) moments of the relative density
fluctuation $\rho_m/\aver{\rho_m}$. This is the ``noise-suppression'' theorem
{}~\cite{bialas1,bialas2}. It assumes that the noise is Poissonian
(cfr.~(\ref{s1:1})) and that the number of counts in all intervals (the total
multiplicity) is unrestricted\footnote{\ If the sum over all intervals of the
number of counts is fixed, a slightly more complicated relation can be
obtained
if the noise has a Bernoulli (multinomial) distribution~\cite{bialas1}}.
The property of Poisson-noise suppression has made measurement of factorial
moments a standard technique, e.g. in quantum optics, to study the statistical
properties of arbitrary electromagnetic fields from photon-counting
distributions. Their utility was first explicitly recognized, for the single
time-interval case, in~\cite{Bedard:67:1} and~\cite{Chang:69} and later
generalized to the multivariate case in~\cite{Cantrell:70}. The authors of
{}~\cite{Chang:69} further stress the advantages of factorial
cumulants compared to factorial moments, since the former measure genuine
correlation patterns, whereas the latter contain additional
large combinatorial terms which may mask the underlying
dynamical correlations (however, see the discussion in Sect.~2.1.2).
Multivariate factorial cumulants are derived from the (natural) logarithm
of the factorial moment generating function.
Taking logarithms of both sides of (\ref{dr:56}), one finds that the
multivariate normalized {factorial cumulants} of the counting
distribution are equal to the multivariate normalized {ordinary cumulants}
of the densities $[\rho \delta]$. This
relation, therefore, extends the noise-suppression theorem to cumulants.
This property is exploited in many fields from quantum
optics~\cite{Cantrell:70} to
radar-physics and astrophysics (see e.g.~\cite{Fry:85}).
\section{Sum-rules}
In an interesting $\alpha$-model analysis of factorial
correlators~\cite{pesch:seixas}, scaling relations are derived between
single-variate and 2-variate factorial moments which are independent of
the dimension of the phase space. The result is stated as follows:
If a correlator $F_{11}(D,\delta)$ is effectively independent of
$\delta$ in a range $\delta<D\leq\delta_0$, then
\begin{equation}
F_{11}(D)=2F_2(2D) -F_2(D).
\label{dr2:1}
\end{equation}
Here, $\delta$ is the interval size and $D$ the distance between the
intervals.
Similar types of relations---or sum-rules---are well-known
in optics since the early 1970's. They are exploited
in so-called Multi-Cathode and Multiple-Aperture Single-Cathode
(MASC) photo-electron counting experiments (see
e.g.~\cite{cantrell:fields,bures} and refs. therein).
Consider again the multivariate multiplicity distribution
$P_M(\XVEC{n}{M})$ giving the joint probability for the
occurrence of $n_1$ particles in a cell $\Omega_1$, $\ldots,$ $n_M$
particles in cell $\Omega_M$, with $\Omega_i\cap\Omega_j=0$,
$\forall i,j$ and $i\ne j$.
Let $n$ be the number of particles counted in the union of the $M$ cells,
\begin{equation}
n=\sum_{m=1}^Mn_m\ \ .
\label{dr2:2}
\end{equation}
The probability distribution of $n$ is given by
\begin{equation}
P(n)=\sum_{n_1=0}^{n}\cdots\sum_{n_M=0}^{n} p_M(\XVEC{n}{M})
\delta_{n,n_1+\cdots+n_M}\, .
\label{dr2:3}
\end{equation}
Define the single-variate factorial moment generating function
\begin{equation}
g(z)=\sum_{n=0}^\infty(1+z)^n\,P(n)\,.
\label{dr2:4}
\end{equation}
The function $g(z)$ can be expressed in
terms of the multivariate generating function
(\ref{dr:51}) as:
\begin{equation}
g(z)=\left.G_M(\XVEC{z}{M})\right|_{z_1=z_2=\cdots=z_M=z}.
\label{dr2:5}
\end{equation}
Equation~(\ref{dr2:5}) allows to express factorial moments of $n$ in terms of
the multivariate factorial moments of $\{\XVEC{n}{M}\}$.
Application of the Leibnitz rule
$$ \left(\frac{\ifmath{{\mathrm{d}}}}{\ifmath{{\mathrm{d}}}\!z}\right)^k f(z)=\sum_{\{a_j\}}\frac{k!}
{a_1\,!a_2\,!\ldots a_k!}
\left(\frac{\ifmath{{\mathrm{d}}}}{\ifmath{{\mathrm{d}}}\,z}\right)^{a_1}\,f_1(z)\cdots
\left(\frac{\ifmath{{\mathrm{d}}}}{\ifmath{{\mathrm{d}}}\,z}\right)^{a_k}\,f_M(z)
$$
to the function
$$f(z)=f_1(z)\cdots f_M(z)$$
leads immediately to the relation
\begin{equation}
\tilde{F}_q=\sum_{\{a_j\}} \tilde{F}^{(M)}_{a_1\ldots a_M} \frac{q!}%
{a_1!\,\ldots\,a_M!}.
\label{dr2:6}
\end{equation}
The summation is over all sets $\{a_j\}$ of non-negative integers such that
$$\sum_{j=1}^{M}a_j=q.$$
Formula (\ref{dr2:6}) may be looked upon as a generalization of
the usual multinomial theorem for factorial moments\footnote{See also
{}~\cite{eggers:phd}}.
Likewise, taking the natural logarithm of
both sides of (\ref{dr2:5}), one obtains an identical
relation as (\ref{dr2:6}) among
single-variate and multivariate factorial cumulants.
As an example, for two rapidity bins ($M=2$) of size $\delta$ separated
by a distance $D$, one finds:
\begin{eqnarray}
\tilde{F}_2 &=& \tilde{F}^{(2)}_{02}
+2\tilde{F}^{(2)}_{11}+
\tilde{F}^{(2)}_{20}\ ,\nonumber\\
\tilde{F}_3 &=& \tilde{F}^{(2)}_{03}
+3(\tilde{F}^{(2)}_{12}+\tilde{F}^{(2)}_{21}) +
\tilde{F}^{(2)}_{30}\ ,\label{fff} \\
\tilde{F}_4 &=& \tilde{F}^{(2)}_{04}
+4(\tilde{F}^{(2)}_{13}+\tilde{F}^{(2)}_{31})
+6\tilde{F}^{(2)}_{22}
+\tilde{F}^{(2)}_{40}\,.\nonumber
\end{eqnarray}
The factorial moments $\tilde{F}_{0i}$ are
determined from the single cell (marginal) counting distribution, whereas
the univariate factorial moments $\tilde{F}_q$ are obtained from the
sum of the counts in the two cells.
The relations derived in \cite{pesch:seixas} follow immediately
from (\ref{fff}) by considering two adjacent cells and normalizing
properly.
Since the derivation of (\ref{dr2:6}) is completely general,
it obviously holds irrespective of the dimension of phase space.
The relations (\ref{fff}) are trivially extended
to more than two cells. They allow to measure high-order
correlators by varying the distances between the cells.
In optics and radar-physics, they are typically used in determining
spatial coherence properties of arbitrary e.m. fields.
\section{Scaling laws}
A major part of this paper is devoted to recent
experimental and theoretical research on possible manifestations
of scale-invariance in high-energy multiparticle production
processes. This work centers around two basic inter-related notions:
intermittency and fractality.
A review of the experimental data accumulated over the last years
will be given in Chapter 4. Theoretical work is discussed in Chapter 5.
In particle physics, intermittency is defined, in a strict sense,
as the scale-invariance of factorial moments
(\ref{dr:44})-(\ref{dr:46})
with respect to changes in the size of phase-space cells (or bins)
say $\delta y$, for small enough $\delta y$:
\begin{equation}
F_{q}(\delta y)\propto (\delta y)^{-\phi_{q}} \;\;\;\;\;\; (\delta y \rightarrow 0).
\label{fq113}
\end{equation}
The power $\phi_q>0$ is a constant at any given (positive integer)
$q$ and called ``intermittency index'' or ``intermittency slope''.
The form of (\ref{fq113}) strictly implies that
the inclusive densities $\rho_q$ and the connected correlation functions
$C_q$ become singular in the limit of infinitesimal
separation ($\delta y\rightarrow0$) in momentum space.
Inspired by the theory of multifractals, scaling behaviour of the
$G$-moments (\ref{gghwa}) has also been looked for in the form
\begin{equation}
G_q (\delta y)\propto (\delta y)^{\tau (q)} \;\;\;\;\;\; (\delta y \rightarrow 0).
\label{gq114}
\end{equation}
To describe the inter-relation of the two proposals, we
briefly discuss the formalism of fractals.
Power-law dependence is typical for fractals~\cite{Man82}, i.e. for
self-similar objects with a non-integer dimension. These range from purely
mathematical ones (the Cantor set, the Koch curve, the Serpinsky gasket
etc.) to real objects of nature (coast-lines, clouds, lungs, polymers etc).
For reviews see~\cite{Zeld87,Pala87,PescTH5891}.
The fractal dimension $D_\ifmath{{\mathrm{F}}}$ is defined as the exponent
which provides a finite
limit
\begin{equation}
0 < \lim_{\epsilon \rightarrow 0} N(\epsilon)\epsilon^{D_{\ifmath{{\mathrm{F}}}}} < \infty
\label{n115}
\end{equation}
for the product of $\epsilon^{D_\ifmath{{\mathrm{F}}}}$ and the minimal number of
hypercubes $N(\epsilon)$
of linear size $l=\epsilon$ (Kolmogorov definition) or $l \leq \epsilon$
(Hausdorff
definition) covering the object when $\epsilon \rightarrow 0$.
To a physicist, the definition becomes more transparent if one considers the
relation between the size $l$ of an object and its mass $M$ as a scaling law:
\begin{equation}
M \propto l^{D_{\ifmath{{\mathrm{F}}}}}. \label{m116}
\end{equation}
For usual objects $D_{\ifmath{{\mathrm{F}}}}$ coincides with the topological
dimension (for a line
$D_{\ifmath{{\mathrm{F}}}}=1$, for a square $D_{\ifmath{{\mathrm{F}}}}=2$ and so on).
The condition $\epsilon \rightarrow 0$ means
in practice that such a law should hold in
some interval of ``rather small'' $\epsilon$-values.
The probability $p_{i}(l)$ to be in a hypercube $N_i(l)$ is proportional
to $l^{D_{\ifmath{{\mathrm{F}}}}}$ at small $l$. Therefore, for a fractal the mean value of the
$q$-th order (ordinary) moment is given by
\begin{equation}
\langle p_{i}^{q}(l) \rangle
\propto l^{qD_{\ifmath{{\mathrm{F}}}}} \;\;\;\; (D_{\ifmath{{\mathrm{F}}}}={\mathrm{const}})\ \ . \label{p117}
\end{equation}
Multifractals generalize the notion of fractals, since for these holds
\begin{equation}
\sum_{i}p_{i}^{q}(l)=\langle p_{i}^{q-1}(l)\rangle \propto l^{\tau (q)} \ \ ,
\label{p118}
\end{equation}
where
\begin{equation}
\tau (q)=(q-1)D_{q}. \label{t119} \end{equation}
The $D_{q}$ are called the R\'enyi dimensions~\cite{Renyi70,Feder88} and
depend
on $q$ (generally, for multifractals they are decreasing functions of $q$).
Sometimes it is more convenient to characterize multifractals by
spectral properties, rather than by their dimensions.
Let us group all the boxes with a singularity $\alpha$ ($p_{i}(l)\sim l^{\alpha},
l\rightarrow 0$) into a subset $S(\alpha )$, where $\alpha $ is called the local mass
dimension. The number of boxes $\ifmath{{\mathrm{d}}} N_{\alpha }(l)$ needed to cover $S(\alpha )$ is
\begin{equation}
\ifmath{{\mathrm{d}}} N_{\alpha }(l)=\ifmath{{\mathrm{d}}}\rho (\alpha )l^{-f(\alpha )}\ \ , \label{dn120}
\end{equation}
where $f(\alpha )$ is the fractal dimension of the set $S(\alpha )$ related to
the R\'enyi dimension. For the sum of moments one obtains:
\begin{equation}
\sum_{i=1}^{N_{i}(l)} p_{i}^{q}(l)\propto \int \ifmath{{\mathrm{d}}}\rho (\alpha )l^{\alpha q-f(\alpha )}\ \ .
\label{dr:121}
\end{equation}
{}From (\ref{dr:121}), one gets by the saddle-point method:
\begin{equation}
D_{q}=\frac{1}{q-1} \mbox{min}_{\alpha}\,\left(\alpha q-f(\alpha )\right)=
\frac{1}{q-1} \left(\bar\alpha q-f (\bar \alpha )\right) \label{dq122}
\end{equation}
with $\bar \alpha $ defined as
\begin{equation}
\left.\frac {\ifmath{{\mathrm{d}}} f}{\ifmath{{\mathrm{d}}}\alpha}\right|_{(\alpha =\bar \alpha )}=q(\bar \alpha ). \label{df123}
\end{equation}
The notion of R\'enyi dimensions $D_q$ generalizes the notion of fractal
dimension $D_{0}=D_{\ifmath{{\mathrm{F}}}}$, information dimension $D_{1}$ and correlation
dimension $D_{2}=\nu$. A R\'enyi dimension, therefore, is often called a
generalized dimension.
The difference between the usual topological dimension $D$ (i.e. the support
dimension) and the R\'enyi dimension is called the anomalous dimension (or
codimension)
\begin{equation}
d_{q}=D-D_{q}. \label{dq124}
\end{equation}
The multifractal method is a widely used tool in many branches of physics
and science in general (cfr.~\cite{Zeld87,Pala87,DremSPU90}).
A direct relation may be established between the exponents of factorial
and generalized moments at
comparatively low values of $q$, much smaller than
effective multiplicities contributing to the sum:
\begin{equation}
\phi _{q}+\tau (q)=(q-1)D. \label{fq125}
\end{equation}
Then the exponents are related to R\'enyi dimension and to codimension as
\begin{equation}
\tau (q)=(q-1)D_q \label{tq126}
\end{equation}
and
\begin{equation}
\phi _q=(q-1)d_q. \label{fq127}
\end{equation}
According to the general theory
{}~\cite {sl,BouGeo}, there exists
``a class of multifractals exhibiting universal properties''.
They are called universal multifractals and
are classified by
a L\'evy index $0 \leq \mu \leq 2$ which allows
the codimension to be expressed as
\begin{equation}
d_q =\frac {C_1}{\mu -1}\cdot \frac {q^{\mu }-q}{q-1} \;\;\;\;\; ( C_1 =
{{\mathrm{const}}} ). \label{dq128}
\end{equation}
The L\'evy index $\mu $ is also known as
the degree of multifractality ($\mu =0$
for mono-fractals). Values $\mu <1$ correspond to
so-called ``calm''
singularities, values $\mu>1$ correspond to ``wild'' singularities.
One can proceed further and try to analyse experimental data at
two different levels of
bin-splitting. For that purpose, it was recently
suggested~\cite{Ratti91,Ratti92} to study
Double Trace Moments (DTM). The procedure
is, first, to sum up $\nu $-th-order moments
of multiplicity distributions at some
bin-splitting level $\Theta $
within bins belonging to a single bin of one
of the previous steps (having bins of size $\Delta$)
and then to calculate their $q$-th moments at that level
\begin{equation}\label{tr}
Tr_{q}^{\nu} \propto \sum_{\Delta}(\sum_{\Theta}n_{m}^{\nu})^{q}
\propto \Delta^{-K(q,\nu)+q-1}.\label{DTM}
\end{equation}
It is claimed~\cite{Ratti91}
that ``the DTM-technique provides a robust estimate of $\mu $
and $C_1$'' for
universal multifractals.
According to
the theory of universal multifractals~\cite{sl,Ratti91},
one should observe the
following factorizable behaviour of ``double'' exponents $ K(q,\nu )$:
\begin{equation}
K(q,\nu )=\nu ^{\mu} K(q,1) \ \ ,\label{fq129}
\end{equation}
where $\mu $ is the same L\'evy index as in (\ref{dq128}).
Experimental results on multifractals and generalized
multifractals, as well as some theoretical implications are
discussed in Subsect.~4.7.7 and in Chapter~5.
\section{Bunching-parameter approach}
A simple mathematical tool alternative to the normalized factorial moments
(2.68-2.70) is the bunching-parameter approach, suggested for high energy
applications in \cite{chek94}. In order to reveal spiky structure of
the events, it is only necessary to study the behaviour of the probability
distribution near the multiplicity $n=q$ by means of the ``bunching
parameters''
\begin{equation}
\eta_q(\delta y) = \frac{q}{q-1}\frac{P_q(\delta y)P_{q-2} (\delta y)}{P^2_{q-1}(\delta y)}\ \ ,
\ \ q>1\ .
\label{BB:128}
\end{equation}
As is the case for the normalized factorial moments, the bunching
parameters $\eta_q$ are independent of $\delta y$ if there are no dynamical
fluctuations. For example, $\eta_q=1$ for all $q$ for the case of
a Poissonian probability distribution.
As the $F_q(\delta y)$, the $P_q(\delta y)$ can be averaged over a number $M$ of
bins. Assuming approximate proportionality of $\bar n_m$ and $\delta y$ at
$\delta y\to 0$ and $P_0(\delta y)\to 1$ for $\delta y\to 0$, one obtains
\begin{eqnarray}
\eta_2(\delta y) & \simeq & F_2(\delta y) \nonumber \\
\eta_q(\delta y) & \simeq & \frac{F_q(\delta y) F_{q-2}(\delta y)}{[F_{q-1}(\delta y)]^2}\ \ ,\
\ q>2
\end{eqnarray}
or
\begin{eqnarray}
\eta_2(\delta y) & \propto & (\delta y)^{-\beta_2} \nonumber \\
\eta_q(\delta y) & \propto & (\delta y)^{-\beta_q}
\end{eqnarray}
with
\begin{eqnarray}
\beta_2 & = & d_2 \nonumber \\
\beta_q & = & d_q(q-1) + d_{q-2}(q-3)-2d_{q-1}(q-2)\ \ ,\ \ q>2\ .
\end{eqnarray}
Expressing $d_q$ in terms of the L\'evy-law approximation (2.125),
\begin{equation}
\beta_q=d_2 \frac{q^\mu + (q-2)^\mu - 2(q-1)^\mu}{2^\mu-2}\ \ \ .
\end{equation}
In case of monofractal behaviour $(\mu=0)$, $\beta_q=0$ for $q>2$. In the
limit of the log-normal approximation $(\mu=2)$, on the other hand,
$\beta_q=d_2$ and all bunching parameters follow the same power-law.
The L\'evy-law approximation allows a simple description of multifractal
properties of random cascade models using only one free parameter $\mu$.
In the bunching-parameter approach, one can make an approximation of the
high-order bunching parameters to obtain a simple {\it linear} expression
for the anomalous fractal dimensions $d_q$, still maintaining the number
of free parameters at one.
Assuming that high-order bunching parameters can be expressed in terms of
the second-order one as
\begin{equation}
\eta_q(\delta y) = [\eta_2(\delta y)]^r\ \ ,\ \ q>2\ \ ,
\end{equation}
the linear expression becomes
\begin{equation}
d_q = d_2 (1-r) + d_2 r \frac{q}{2}\ \ .
\label{BB:134}
\end{equation}
The use of bunching parameters is interesting, because it gives a general
answer to the problem of finding a multiplicity distribution leading
to intermittency: according to (\ref{BB:128}), any multiplicity distribution
can be expressed as
\begin{equation}
P_q(\delta y) = P_0 (\delta y) \frac{ [P_1(\delta y)/P_0(\delta y)]^q}{q!} \prod^q_{\ell=2}
[\eta_\ell(\delta y)]^{q+1-\ell}, \ \ \ q>1\ \ .
\end{equation}
The possible forms of multiplicity distributions with multifractal
behaviour of $d_q$ (\ref{dq124}) are discussed in~\cite{chek95,CheKu}.
\section{The wavelet transform}
An increase of factorial and cumulant moments with decreasing
bin sizes reflects
a widening of a multiplicity distribution, i.e. an increase of multiplicity
fluctuations in individual events. This phenomenon can be studied by other
methods, as well. In particular, the so-called wavelet transform seems to
be suited for that purpose.
The wavelet transform is of particular importance in pattern
recognition.
This is a more general problem than the fluctuation study
itself, since it involves the analysis of individual event shapes, not
only the event ensemble, and may become of interest in the analysis of
very-high multiplicity events.
It is shown~\cite{ADREP} that, for pattern recognition, the wavelet
transform is about two orders of magnitude more efficient than
ordinary Fourier analysis.
An application of wavelets to multiparticle production processes
has been proposed in~\cite{CGL}.
The main principle of the wavelet transform is to study the dependence of
fluctuations on the phase-space bin size by the so-called difference method.
One considers the difference between the histogram of an individual
event at a definite resolution to the corresponding histogram at a (e.g.,
twice)
finer resolution. Proceeding step by step, one is able to restore the whole
pattern of fluctuations.
Let us explain how this procedure can be applied to an individual event. We
consider the one-dimensional projection of the event onto the rapidity
interval $\Delta Y$. Any $n$-particle event can be represented by the
histograms of particle densities $\rho = dn/dy$ at various resolutions. The
simplest information is obtained from the value of the average density
$\langle \rho \rangle = n/\Delta Y$. To consider the forward-backward
correlations, one splits the rapidity interval $\Delta Y$ into two equal parts
and gets the forward and backward average densities $\langle \rho _{f,b}
\rangle = 2n_{f,b}/\Delta Y$, where $n_{f,b}$ are the forward (backward)
multiplicities with $n_f + n_b =n$. Proceeding further to the $J$-th step,
we approximate the event in terms of the histogram with $2^J$ bins.
Let us construct now the difference of the two histograms described above.
Namely, we subtract the average density from the forward-backward histogram
and get another histogram with positive ordinate at one side and negative at
the other, demonstrating the forward-backward fluctuations in the event.
Splitting the forward and backward regions further into equal halves,
one gets the histogram at $J=2$. Its difference from the forward-backward
histogram at $J=1$ reveals the fluctuations at finer resolution.
Iterating
to higher values of $J$, one studies how fluctuations evolve
at ever finer resolution. The set of difference histograms is called the
wavelet transform of the event. The above procedure
corresponds to the so-called Haar-wavelet transform. Those interested in
mathematical details are referred to~\cite{DMK}.
The wavelet transform provides direct information on the
evolution of fluctuations at different scales, i.e. on the dynamics of
individual high-multiplicity events revealing their clustering (and
subclustering) structure. A generalization to factorial (and cumulant)
wavelets is possible~\cite{CGL}. The simplest cascade models show such
remarkable properties of wavelet transforms~\cite{CGL} as
(quasi)diagonalization of their correlation density matrices, scaling
exponents etc. It is interesting to note that the
equations for the generating functions of wavelet transforms~\cite{CGL} look
very similar to the ``gain-loss'' equations (in particular, to QCD equations)
discussed at the end of Chapter~5. All those features are yet
to be studied.
The very first application to experimental data is presented in
{}~\cite{Suzu95}, where wavelet spectra of JACEE events are
studied.
\chapter{Experimental survey on correlations}
In this chapter, we review experimental results on ``classical''
correlations, a subject with a long history in particle physics.
It was instrumental in establishing fundamental concepts
of hadrodynamics, such as short-range order, which are an
essential ingredient of all popular Monte-Carlo models of
hadronization. With the exception of Bose-Einstein interferometry,
the field lay dormant for several years, but was revived with the
introduction of generalized concepts. The data
cover a variety of multiparticle-production processes ranging
from $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ annihilation to nucleus-nucleus collisions.
In Chapter~4, we shall review material on factorial moments and related
quantities, obtained since 1986. At that time, a pioneering suggestion
was made to investigate the patterns of particle density fluctuations
in multihadronic events: the intermittency idea. Measurement of
factorial moments opened a way to establish possible scale-invariance
and fractal behaviour in hadrodynamics.
Interest in correlation functions received a vigorous boost
when their intimate connection with factorial moments was
realized (see Chapter~2).
Both are now explored in parallel with novel techniques.
These offer promising perspectives towards a long overdue
unified approach to correlation phenomena, including
Bose-Einstein interferometry.
Another obviously related subject, the phenomenology of
multiplicity distributions \cite{CarShih}, is not explicitly
covered here.
Multiplicity distributions inspired many early
ideas on scale-invariance and phase-transition analogies
in multiparticle production, such as Koba-Nielsen-Olesen scaling~\cite{KNO}
and the Feynman-Wilson liquid picture~\cite{FWfluid}.
However, the major part of the data relate either to full phase space
or to sizable portions of it.
It remains an interesting task for the future to
explain the ``large-scale'' properties of multiplicity
distributions in terms of correlation function behaviour
at ``small distances'', the main subject of this paper.
Of course, the factorial moments discussed in
Chapter~4 are just another representation
of multiplicity distributions and their increase with decreasing bin
size reveals the evolution of the multiplicity distribution.
\section{Rapidity correlations}\\
The study of correlation effects in particle production processes provides
information on hadronic production dynamics beyond that obtained from
single-particle inclusive spectra. Correlations in rapidity $y$, as defined in
Sect.~2.1, have been studied in various experiments on $\re^+\re^-$, lepton-nucleon,
hadron-hadron, hadron-nucleus and nucleus-nucleus collisions. Strong
$y$-correlations have been observed in all experiments in one form or another,
depending on the specific form of the correlation function, type of
interaction, kind of particles, the kinematic region under consideration,
etc. The main conclusions were (for early reviews see~\cite{Foa75,Whit76}):
\begin{itemize}
\itemsep=-2mm
\item[1.] Two-particle correlations are strong at small interparticle
rapidity-distances $|y_1-y_2|$ (see Fig.~3.1).
\item[2.] They strongly depend on the two-particle charge
combination.
\end{itemize}
Rapidity correlations are now being studied with renewed attention.
One reason is that their structure at very small rapidity distances is
directly related to self-similar particle-density fluctuations
(intermittency), a topic to be covered in Chapter~4.
\subsection{Correlations in hadron-hadron collisions}\\
In Fig.~3.2 the pseudo-rapidity correlation function $C_2(\eta_1,\eta_2)$
as defined in (\ref{dr:ex1}) is given for
$\eta_1=0$, as a function of $\eta_2=\eta$,
for the energy range between 63 and 900 GeV~\cite{Anso88}. Whereas
$C_2(0,\eta)$ depends on energy, the short-range correlation $C_\ifmath{{\mathrm{S}}}$
defined in (\ref{dr2.9}) does not strongly depend on energy and has a full
width of about 2 units in pseudo-rapidity.
The function $C_\ifmath{{\mathrm{L}}}$ is not a two-particle correlation, but derives from
the difference in the single particle distribution function for
different multiplicities. As can be seen in Fig.~3.2b, $C_\ifmath{{\mathrm{L}}}$ is considerably
wider than $C_\ifmath{{\mathrm{S}}}$ and
increases with energy (the 63 GeV data are from~\cite{Amen76}).
In Fig.~3.3, the semi-inclusive correlation $C^{(n)}_2(\eta_1,\eta_2)$ for
$\Pp\Pap$
collisions at 900 GeV~\cite{Fugl87} is compared to the UA5
Cluster Monte Carlo (MC) GENCL~\cite{Alner87}, as well as to the FRITIOF~2
{}~\cite{FRIT} and PYTHIA~\cite{PYTHIA} Monte Carlos, for charge
multiplicity $34\leq n\leq 38$. The Cluster MC is designed to fit just these
short-range correlations, but also FRITIOF~2 is doing surprisingly well
(see however Subsect.~4.4.4).
At lower energy, the NA23 Collaboration~\cite{Bail88} has studied the
short-range correlation of charged particles in pp collisions of $\sqrt s=26$
GeV in terms of $K_2(y_1,y_2)$ defined in (\ref{dr:2.7}). Only events with
charge multiplicity $n>6$ are used. The positive short-range correlations are
in agreement with those found earlier at $\sqrt s= 53$ GeV~\cite{Break82}.
The NA23 data are compared to single-string LUND~\cite{LUND} and to a
two-chain Dual-Parton Model (DPM)~\cite{DPM} in Fig.~3.4.
The one-string model (without gluon radiation) does not at all describe
the short-range rapidity correlation in the data. The two-chain model
does better, but remains unsatisfactory.
Somewhat better but still insufficient agreement is obtained
by renormalizing the MC events
to the experimental multiplicity distribution (not shown). The effect of
Bose-Einstein correlations in the (++) and (--~--) data is found
to be insignificant, as may be expected for data integrated over
transverse momentum $p_\ifmath{{\mathrm{T}}}$ and azimuthal angle $\varphi$. Obviously, more
chains, possibly with higher $p_\ifmath{{\mathrm{T}}}$, are needed to explain short-range
order with fragmentation models, even below $\sqrt s\approx30$ GeV.
NA22 results for $C_2(0,y_2)$ and $\tilde C_2(0,y_2)$
(Eqs.~\ref{dr:ex1}, \ref{dr:2.5b}) for $\pi^+$p
and K$^+$p collisions at $\sqrt s$=22 GeV~\cite{Aiva91}
are compared with FRITIOF~2, a 2-string DPM
and QGSM~\cite{QGSM} predictions in Fig.~3.5a,b. FRITIOF
and 2-string DPM largely underestimate the correlation.
QGSM reproduces $C^{--}_2(0,y_2)$ very well and even overestimates
$C^{++}_2(0,y_2)$ and $C^{+-}_2(0,y_2)$.
It has been verified that the differences between
QGSM and FRITIOF or DPM
are not due to the different treatment of tensor mesons
(only included in the latter two).
In Fig.~3.5c, FRITIOF and QGSM are compared to the NA22 data
in terms of the short-range contribution $\tilde C_{\ifmath{{\mathrm{S}}}}(0,y_2)$.
The $(+-)$ short-range correlation is
reproduced reasonably well by these models.
For equal charges, however, the strong
anti-correlation predicted by FRITIOF is
not seen in the data. QGSM contains a small equal-charge correlation due to
a cluster component, but still underestimates its size.
Similar discrepancies are also observed
in semi-inclusive (fixed multiplicity) data for each charge
combination (not shown here). They are even larger
than in the inclusive data, also in the QGSM model.
{}From this brief survey, we conclude that
in hadron-hadron collisions two-particle correlations are
badly reproduced and generally underestimated in currently used models.
\subsection{Correlations in $\protect\re^+\re^-$ and $\protect\mu^+p$-collisions}\\
Fig.~3.6 shows $K^{+-}_2(y_1,y_2)$ and $K^{--}_2(y_1,y_2)$ for muon-nucleon
interactions at 280 GeV/$c$ \cite{Arne86}. A steep peak is seen at
$y_1=y_2=0$ for $K^{+-}_2$, with two shoulders along the diagonal $y_1=y_2$.
On the other hand, $K^{--}_2$ is below 0 for most of the distribution, but
we shall see that the most impressive correlation is in fact coming from
$y_1\approx y_2$, just for this case. As in hadron-hadron collisions,
correlations are strong and depend on the two-particle charge combination.
Fig.~3.7 shows $K_2(y_1,y_2)$ in $\mu^+$p interactions at 280 GeV/$c$ with
$y_1$ $\epsilon\ [-0.5,0.5]$, the hadronic invariant mass $W$ in the interval
$13<W<20$ GeV and for $n\geq 3$~\cite{Male90Fig88}, together with the NA22
non-single-diffractive M$^+$p sample, $n\geq 2$~\cite{Aiva91}. Correlations
in $\mu^+$p seem smaller than in NA22, but one has to consider a possible
energy dependence. Indeed, extrapolating from the energy dependence of
$K_2(0,0)$ published in~\cite{Male90Fig88}, one finds quite
similar values for $\mu^+$p at 22 GeV and M$^+$p in NA22.
In Fig.~3.8a,b we compare the function $\tilde K_2(0,y)$ for the NA22
non-single-diffrac\-tive M$^+$p sample (charge multiplicity
$n\geq 2$)~\cite{Aiva91} with that for $\re^+\re^-$-annihilation at the same
energy ($\sqrt s$=22 GeV)~\cite{Alth85Chwas88}. The values of
$\tilde K_2(0,y)$ are larger for $(++)$ pairs than for $(--)$ in
meson-proton (M$^+$p) reactions; for $(--)$ and $(+-)$ pairs they agree
with $\tilde K_2$ for $\re^+\re^-$ annihilation in the central region.
A comparison of the correlation functions for $\re^+\re^-$-annihilation and
non-single-diffract\-ive M$^+$p collisions throughout the full kinematic
region with $y_1$ $\epsilon\ [-1,0]$ is shown in Fig.~3.8c for charged pairs.
The $\re^+\re^-$ data are given at $\sqrt s=14$ and 44 GeV~\cite{Alth85Chwas88}.
At $y_2=y_1$, the 22 GeV M$^+$p correlation lies between the $\re^+\re^-$ results.
The shape is, surprisingly, more symmetric than in $\re^+\re^-$.
For $\mu^+\Pp$ ~\cite{Arne86,Male90Fig88} and $\re^+\re^-$ collisions
\cite{Alth85Chwas88,Podo91,Act92}, the LUND-type Monte Carlo is reported
to reproduce the majority of the experimental distributions. In~\cite{Bail88}
it is shown that this is mainly due to the inclusion of hard and soft gluon
effects. However, important underestimates of $K_2(y_1,y_2)$ are still
observable, in particular in the central and current fragmentation regions.
For $\re^+\re^-$~\cite{Podo91}, this is shown in Fig.~3.9, where $K_2(y_1,y_2)$ is
compared to the LUND model (JETSET 7.2 PS) as a function of $y_1-y_2$ (dotted
line), for the full sample (upper plots) and for a two-jet sample (lower
plots). In all cases, the LUND model underestimates the correlation at
$y_1-y_2=0$. In general, the disagreement becomes smaller
when Bose-Einstein correlations are included (full lines).
The main feature to note is that correlations are much
weaker in the two-jet sample than in the full sample.
Furthermore, correlations are larger for $y\leq 0$ (left
plot), i.e. in the hemisphere opposite the most energetic jet,
than for $y>0$ (right plot). These two observations, again, point to
hard gluon radiation as the main source of two-particle correlation in
$\re^+\re^-$ collisions.
A systematic test of analytic QCD calculations and of QCD
Monte-Carlo models for two-particle correlations
has been performed by OPAL~\cite{Act92}. The authors study the function
\begin{equation}
R(\xi_1,\xi_2) = K_2(\xi_1,\xi_2)+1
\end{equation}
with $\xi=\ln(1/x_\ifmath{{\mathrm{F}}})$, $x_\ifmath{{\mathrm{F}}}=2p/E_{{\mathrm{cm}}}$ being the Feynman variable,
i.e. the particle momentum $p$ in the cms normalized to half the
cms energy $E_{{\mathrm{cm}}}$. In Fig.~3.10, $R$ is plotted as a function of
$(\xi_1-\xi_2)$ for $(\xi_1+\xi_2)$ centered at the values 6, 7 and 8,
respectively. Fig.~3.10a proves that a next-to-leading order
calculation~\cite{Fo91} (full lines) is better than leading order (dashed),
but still overestimates the overall level of the correlation for
any reasonable value of $\Lambda$. Since the next-to-leading correction
is large, still higher-order terms are needed. It is therefore likely
that a satisfactory analytical treatment of correlations, even at the
parton level, will not be obtained in the very near future.
Higher-order effects are, in an average sense, included in the
existing Monte-Carlo models. In Fig.~3.10b, the same data are compared
to the coherent parton shower models JETSET~PS~\cite{LUND},
HERWIG~\cite{Mar88} and ARIADNE~\cite{Pet88}. The latter gives an
excellent fit to the data, JETSET lies slightly below (within
uncertainty of parameters), but HERWIG considerably above. The agreement
of JETSET could only slightly be improved by including Bose-Einstein
correlations. As far as incoherent parton shower models
are concerned, none of the various versions of COJETS~\cite{Odo84} gives
a particularly good representation of the correlation data.
All the models were tuned on the OPAL data in terms of event shapes and
generally describe single-particle distributions. It is clear that
correlations
allow better and more discriminative tests than more integrated quantities.
We have mentioned the difficulties string-hadronization models experience
in predicting like-sign correlations in hadron-hadron collisions. It
is important to verify if the otherwise successful $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ models
are also able to reproduce correlations between charge-separated
systems such as $(+-)$ and $(\pm\pm)$ particle pairs.
\subsection{Charge dependence}
How $C_\ifmath{{\mathrm{S}}}$ and $\tilde C_\ifmath{{\mathrm{S}}}$ depend on the charge of the pairs
is shown in Fig.~3.11 for the combinations $(--),(++)$ and $(+-)$ in
NA22~\cite{Aiva91}. The short-range correlation is
significantly larger for $(+-)$ than for $(--)$
and $(++)$ combinations. This is also seen in the EMC data~\cite{Arne86}.
Resonance production is a likely explanation of this
difference. For like charges, a small enhancement is seen near
$y_1\approx y_2\approx 0$ above a large negative background. This is
possibly due to Bose-Einstein interference.
\subsection{Charge-multiplicity dependence}
The multiplicity dependence of $\tilde C_2^{(n)}(0,y)$ for the $(+-)$
combination is shown in Fig.~3.12 \cite{Aiva91}. Near the maximum at
$y=0$ the correlation function is approximately Gaussian and narrows
with increasing $n$. In Fig.~3.13a are presented the values of
$\tilde C_2^{(n)}(0,0)$ as a function of $n$ for three charge combinations.
Within errors, $\tilde C_2^{(n)}(0,0)$ is independent of $n$, but
consistently higher for $(+-)$ and $(--)$ than for $(++)$. The reason for
the difference between $(--)$ and $(++)$ probably lies in the positive charge
of both beam and target.
On the other hand, an increase of $\tilde C_2^{(n)}(|\eta_1-\eta_2|)$ with
$1/(n-1)$ is found~\cite{EGGE75} when averaging over a region $|\eta|<2$
(Fig.~3.13b). Since $\tilde C_2$ becomes smaller when moving away from the
center, and that may happen faster for higher than for lower $n$, this is not
necessarily in contradiction with the data in Fig.~3.13a.
\subsection{Transverse momentum dependence}
The search for density fluctuations, described in later sections, has
revealed the importance of correlations in multi-dimensional phase space.
It is, therefore, of interest to gain insight into the transverse momentum
($p_\ifmath{{\mathrm{T}}}$) dependence of rapidity correlations. Early results on this topic
can be found in~\cite{Biswas76}.
Recent data on $K_2(0,y_2)$~\cite{NA22} for all particles and for particles
with $p_\ifmath{{\mathrm{T}}}$ smaller or larger than 0.3 GeV/$c$, plotted in Fig.~3.14,
indeed reveal a strong sensitivity to transverse momentum.
The correlation function is largest,
and stronger peaked, near $y_2=0$ for $p_\ifmath{{\mathrm{T}}}<0.3$ GeV/$c$, in particular for
$(--)$-pairs. A similar effect was noted already in~\cite{Biswas76}.
The data of Fig.~3.14 were fitted with the functions
\begin{eqnarray}
f_1 &=& c \exp [-(y-y_0)^2/2\sigma^2]\ \ \ \ \hfil{\mbox{(full\ line)}}\ \ , \\
f_2 &=& a \exp(-b|y|)\ \ \ \ {\mbox{(dashed)}} \ \ ,
\end{eqnarray}
with $c$, $y_0$, $\sigma$, $a$ and $b$ as free parameters. Even though
for low $p_\ifmath{{\mathrm{T}}}$ the data point at $y_2=0$ lies systematically above the curve,
$K_2(0,y_2)$ is well fitted by the Gaussian $f_1$ but not by the exponential
$f_2$, in this one-dimensional projection on rapidity.
Changing to the variables $x_1=(y_2+y_1)/2$ and $x_2=(y_2-y_1)/2$, a
steepening is observed at small $x_2$ (not shown). For like-charge pairs,
this becomes particularly sharp when the bin size is
reduced to $\delta x_2=0.1$. For the latter, $C_2(x_1,x_2=0)$ increases and
both a Gaussian and an exponential can fit the correlation function.
\subsection{Strange particles}
In string-fragmentation models, first-rank hadrons are formed from
neighbouring quark-antiquark pairs tunnelling out of the vacuum. The
hadronic final states, therefore, show short-range order due to local
flavour conservation. Using stable mesons only, this characteristic
property is difficult to study experimentally because of the large $\Pq\Paq$
combinatorial background. What is needed is a flag identifying the $\Pq\Paq$
pairs created together. A suitable choice is strangeness since the number
of $\ifmath{{\mathrm{s}}}\overline{\rs}$-pairs per event is small and the combinatorial background
strongly reduced.
Good strangeness identification is available for $\re^+\re^-$ annihilation
in the TPC detector at $\sqrt s=29$ GeV~\cite{Aiha84}. This collaboration
observes significant short-range $\PK^+\PK^-$ correlations in $y$,
well reproduced by the LUND model and by the Webber QCD model.
In hadron-hadron collisions, strange particle pairs have been studied by
the NA23 Collaboration~\cite{Asai87}. The distribution in the rapidity
difference $\Delta y$ for two $\PK^0$'s is given in Fig.~3.15a, for a $\PK^0$ and
a $\Lambda^0$ in Fig.~3.15c. The results are compared to the single-string LUND
model. As is the case for non-strange particles, the model slightly
underestimates the rapidity correlation.
\section{Azimuthal correlations}
In interactions of unpolarised particles, no distinguished direction
exists in the plane transverse to the beam and the distribution in the
azimuthal angle $\varphi$ is uniform. Still, a two-particle correlation exists
also in $\varphi$ and is visible in the distribution $W(\Delta\varphi)$ of
$\Delta\varphi=|\varphi_1-\varphi_2|$, the azimuthal angle between two particles,
$\Delta\varphi\in(0,\pi)$. The azimuthal correlation may depend on the charge of
the particles in the pair, on the rapidity distance $\Delta y = |y_1-y_2|$
between these particles and on their transverse momentum.
The first experiments to extensively study two-particle correlations as a
function of both rapidity and azimuthal angular separation~\cite{EGGE75,OH}
already showed that the correlation at small rapidity distance is strongest
when the two particles are produced in the same or opposite directions in
transverse momentum (see Fig.~3.16). The correlation-length in rapidity
is larger towards $\Delta\varphi=\pi$ than towards $\Delta\varphi=0$. Furthermore, significant
differences in the shape of the joint rapidity and azimuthal correlation
functions have been observed for pairs of like and unlike pions~\cite{OH}.
\vspace{2mm}
In Fig.~3.17, the distribution $W(\Delta\varphi,\Delta y)$, normalized to unity, is
shown as a function of $\Delta\varphi,$ for all charge combinations, in the intervals
$\Delta y<1, ~1<\Delta y<2$ and $2<\Delta y<3$~\cite{NA22}. A {horizontal line} at the
average value $1/\pi$ corresponds to a flat distribution in $\Delta\varphi$. The
distribution is influenced by conservation of transverse momentum, by the
decay of resonances (mainly for unlike-sign particles)
and by Bose-Einstein correlations (for like-sign particles).
In all cases, $W$ is larger than $1/\pi$ for $\Delta\varphi>\pi/2$ and has a maximum
at $\Delta\varphi = \pi$. Except for $(--)$ pairs at $\Delta y<1$, the $W$ function is
smaller than $1/\pi$ for $\Delta\varphi < \pi /2$. Such a global
anti-correlation follows from transverse momentum conservation.
Model predictions are shown in Fig.~3.17 for FRITIOF~2 (dot-dashed),
two-string
DPM (full) and multi-string QGSM (dashed). The comparison with the data shows
that it is much easier to account for
azimuthal correlations at large than at small
$\Delta y$. At small $\Delta y$ the models differ from each other and from the
experimental data. The QGSM shows somewhat better agreement with
experiment than the other models. This is a consequence of the
multi-string structure of QGSM, where strong azimuthal correlations in a
single string are destroyed, with the result that the $\Delta\varphi$-dependence
is weaker than in two-string models.
Differences between experiment and all models exist at small $\Delta\varphi$ and
$\Delta y<1$, in particular for $(--)$ pairs. Bose-Einstein correlations, not
included in the models, may explain this disagreement. The influence of
Bose-Einstein correlation can also be observed in the $(++)$ combination, but
is smaller because of the influence of the (positive) beam particle.
Azimuthal distributions are shown in Fig.~3.18 for particles with
$\Delta y<1$, for
$p_\ifmath{{\mathrm{T}}}<0.30$ GeV/$c$ and for $p_\ifmath{{\mathrm{T}}}>0.30$ GeV/$c$, together with
model calculations. A comparison of these figures reveals that azimuthal
correlations have a strong $p_\ifmath{{\mathrm{T}}}$-dependence. Large {\it positive}
azimuthal correlations exist at small $\Delta\varphi$ and $\Delta y<1$ for like-sign
particles with small $p_\ifmath{{\mathrm{T}}}$. As the transverse momentum of particles
increases, the peak at $\Delta\varphi = \pi$ becomes more pronounced. This is
reproduced by the models and reflects momentum conservation.
For $\Lambda\bar \Lambda$ pairs, an azimuthal correlation has been observed in
MARK II at 29 GeV~\cite{Vais85}. Similar $\PK^+\PK^-$ correlations are seen
in the exclusive hh final state $\PK^-\Pp\to$\break
$\Pp\PK^+\PK^-\PK^-\pi^+\pi^-$ at 32 GeV/$c$~\cite{MaZP86}.
Azimuthal correlations between $(+ -)$ and $(++,--)$ charge combinations have
been studied in $\mu$p collisions~\cite{Arne86} for $|\Delta y|<1$ and $|\Delta y|>1$.
The distribution $W(\Delta\varphi)$ is described fairly well by the LUND model
including primordial $k_\ifmath{{\mathrm{T}}}$ and gluons, except that for $|\Delta y|<1$ it
slightly underestimates the anti-correlation for $(+ -)$ and overestimates
it for $(++,--)$.
In the azimuthal correlation of $\PK^0$ pairs (Fig.~3.15b) and of $\PK^0\Lambda^0$
(Fig.~3.15d) studied by NA23~\cite{Asai87}, the data tend to show pairs of
small $\Delta\varphi$ not pre\-sent in low-$p_\ifmath{{\mathrm{T}}}$ LUND (solid line).
By the same collaboration, the azimuthal correlation is studied~\cite{Bail88}
in terms of the asymmetry parameter
\begin{equation}
B=[N(\Delta\varphi >\pi/2) - N(\Delta\varphi<\pi/2)]/N_{{\mathrm{all}}}
\end{equation}
for hadron pairs with
\noindent
a) opposite charge ($\ifmath{{\mathrm{h}}}^+\ifmath{{\mathrm{h}}}^-$)\\
b) equal charge ($\ifmath{{\mathrm{h}}}^+\ifmath{{\mathrm{h}}}^++\ifmath{{\mathrm{h}}}^-\ifmath{{\mathrm{h}}}^-$)\\
c) possibly opposite strangeness ($\Lambda^0\ifmath{{\mathrm{h}}}^+,x_\Lambda<-0.2$)\\
d) no opposite strangeness ($\Lambda^0\ifmath{{\mathrm{h}}}^-,x_\Lambda<-0.2$),
\noindent
for $\Delta y<2$ and for $\Delta y>2$. No azimuthal correlation is seen for $\Delta y>2$
in all cases and for $\Delta y<2$ in case of no common $\Pq\Paq$ pairs
($\ifmath{{\mathrm{h}}}^+\ifmath{{\mathrm{h}}}^++\ifmath{{\mathrm{h}}}^-\ifmath{{\mathrm{h}}}^-,\Lambda^0\ifmath{{\mathrm{h}}}^-$). For $\ifmath{{\mathrm{h}}}^+\ifmath{{\mathrm{h}}}^-$ and $\Lambda^0\ifmath{{\mathrm{h}}}^+$, the
parameter $B$ is compared to low-$p_\ifmath{{\mathrm{T}}}$ LUND and DPM predictions in
Table~3.1.
\vspace{1mm}
\begin{center}
{\bf Table 3.1} Asymmetry parameter $B$
\vskip 4mm
\begin{tabular}{llll}
\hline
&Experiment &LUND &DPM \\
\hline
$\Lambda^0\ifmath{{\mathrm{h}}}^+ (\Delta y<2)$ & 0.18~$\pm$0.03& 0.30~$\pm$0.01& 0.19~$\pm$0.01 \\
$\ifmath{{\mathrm{h}}}^+\ifmath{{\mathrm{h}}}^- (\Delta y<2)$& 0.066$\pm$0.003& 0.126$\pm$0.002& 0.106$\pm$0.002 \\
\hline
\end{tabular}
\end{center}
\vspace{2mm}
\noindent
The parameter $B$ is strongly overestimated in single-string low-$p_\ifmath{{\mathrm{T}}}$
LUND and still too large in the two-string DPM. Furthermore, $B$ increases
with the sum of the transverse momenta (Fig.~3.19) but less strongly than in
the models.
The azimuthal correlation has also been studied for $\ifmath{{\mathrm{c}}}\bar \ifmath{{\mathrm{c}}}$ pairs in
$\ifmath{{\mathrm{D}}}\bar\ifmath{{\mathrm{D}}}$ production. An asymmetry is indeed observed in $\pi^-\Pp$
collisions
at 360 GeV/$c$~\cite{Agui85}. Also there, the LUND model overestimates the
effect.
As shown on $\pi^-$N interactions at $\sqrt s=26$ GeV \cite{Adamo95}, also
NLO perturbative QCD calculations overestimate the azimuthal asymmetry for
$\ifmath{{\mathrm{D}}}\bar \ifmath{{\mathrm{D}}}$ pairs (Fig.~3.20a). Agreement can be obtained with a
model \cite{Frixi94} where a (Gaussian shaped) transverse component is added
to the incoming parton momentum before performing the NLO perturbative
QCD calculation (Fig.~3.20b).
\section{Correlations on the parton level}
The OPAL collaboration \cite{Act93} has compared hadronic azimuthal
correlations to coherent and incoherent shower models (Fig.~3.21). The
coherent models JETSET~PS with angular ordering \cite{LUND}, HERWIG
\cite{Mar88} and ARIADNE \cite{Pet88} describe the azimuthal correlations
in hadronic $\PZz$ decays, but the incoherent models JETSET~PS without
angular ordering \cite{LUND} and COJETS \cite{Odo84} fail for
$\varphi\ ^>\hs-2.5mm_\sim\ \pi/2$.
The hadronization of a quark-antiquark pair at high virtuality is
currently thought to proceed via parton showering \cite{Dok92}.
QCD implies that this parton showering be coherent. The
coherence can be incorporated into Monte-Carlo programs as angular
ordering\cite{ao}, whereby for each successive branching the gluon is
emitted at a smaller angle.
Furthermore, the idea of local parton-hadron duality (LPHD)\cite{lphd}
suggests that features at the parton level survive the fragmentation process.
We can, therefore, expect that the coherence of the parton radiation
will be reflected in angular ordering of the observed particles.
As a method particularly sensitive to angular ordering,
particle-particle correlations PPC and their
asymmetry PPCA \cite{aleph_note,Syed94} are examined in a way analogous
to the study of energy-energy correlations \cite{basham},
\begin{eqnarray}
\ifmath{\mathrm{PPC}} (\chi)&=& \frac{1}{\Delta\chi}
\langle 2 \sum^n_{i<j}
\frac{1}{n^2}
\delta_{\mathrm{bin}}(\chi-\chi_{ij})\rangle
\\
\ifmath{\mathrm{PPCA}} (\chi) &=& \ifmath{\mathrm{PPC}} (180^{\circ}-\chi)-\ifmath{\mathrm{PPC}} (\chi)\ \ ,
\end{eqnarray}
where $\chi_{ij}$ is the full spatial angle between tracks $i$ and $j$,
$\langle~\rangle$ is the average over all events in the sample, $n$ is the
number of charged tracks in an event, and $\Delta\chi$ is the bin
width. The function $\delta_{\mathrm{bin}}(\chi-\chi_{ij})$ is 1 if
$\chi_{ij}$ and $\chi$ are in the same bin and 0 otherwise.
At $\sqrt{s}=M_\ifmath{{\mathrm{Z}}}$, the fraction of two-jet events is very high. For
two-jet events, particles in different jets will in general be separated
by an angle $\chi$ greater than $90^\circ$. The PPC for $\chi > 90^\circ$
can, therefore, serve as an indication of what the PPC
{\em within}\/ a jet ($\chi < 90^\circ$) would be
{\em in the absence}\/ of angular ordering. By forming the asymmetry,
these `uninteresting' correlations are effectively subtracted. The effects
of angular ordering should, therefore, be more directly observable in the
PPCA than in the PPC. Note, however, that the sign convention following
\cite{basham} leads to a {\em negative} sign for a {\em positive} correlation.
Figures 3.22a and b show the PPCA distribution of L3 data (corrected for
detector effects \cite{Syed94}) compared to coherent and incoherent
Monte-Carlo models, respectively.
In Fig.~3.22b we see that for $\chi\ ^<\hs-2.5mm_\sim\ 60^0$ JETSET~7.3 ~PS without angular
ordering (incoherent) disagrees strongly with the data, while being in fair
agreement at larger values of $\chi$. COJETS is seen not to reproduce the data
over the entire angular range. On the other hand, in Fig.~3.22a, the coherent
Monte Carlo models, JETSET with angular ordering, HERWIG, and ARIADNE
all reproduce the data reasonably well over the full angular range. Note
that the disagreement of the incoherent models can not be due to the
Bose-Einstein effect. Turning this effect off in the non-angular ordered
JETSET model does not raise but lower its PPCA points.
So, the data from the L3 experiment strongly disfavour the incoherent models.
\section{Three-particle rapidity correlations}\\
Whether dynamical correlations exist beyond the two-particle correlations
discussed so far is of crucial importance for much of the present search for
scaling phenomena in multiparticle processes, a subject treated in Chapter~4.
With conventional techniques, this question is not easy to answer
and beyond the sensitivity of many experiments.
Nevertheless, three-particle
correlations in rapidity have been looked for in a number of experiments
{}~\cite{Whit76,Aiva91,Buma79,Azim80,BreakEP88}.
The third order normalized factorial cumulant is defined as
[cfr.~(\ref{a:4b})]:
\begin{equation}
K_3(y_1,y_2,y_3) = C_3(y_1,y_2,y_3)/\frac{1}{\sigma^3_{{\mathrm{inel}}}} \frac{\ifmath{{\mathrm{d}}}\sigma}
{\ifmath{{\mathrm{d}}} y_1} \frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_2} \frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_3}\ \ ,
\end{equation}
\begin{equation}
C_3(y_1,y_2,y_3)=\frac{1}{\sigma_{{\mathrm{inel}}}} \frac{\ifmath{{\mathrm{d}}}^3\sigma}{\ifmath{{\mathrm{d}}} y_1 \ifmath{{\mathrm{d}}} y_2 \ifmath{{\mathrm{d}}} y_3}
+ 2 \frac{1}{\sigma^3_{{\mathrm{inel}}}} \frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_1} \frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_2}
\frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_3} -
\end{equation}
\begin{eqnarray*}
- \frac{1}{\sigma^2_{{\mathrm{inel}}}} \frac{\ifmath{{\mathrm{d}}}^2\sigma}{\ifmath{{\mathrm{d}}} y_1\ifmath{{\mathrm{d}}} y_2}
\frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_3} -
\frac{1}{\sigma^2_{{\mathrm{inel}}}} \frac{\ifmath{{\mathrm{d}}}^2\sigma}{\ifmath{{\mathrm{d}}} y_2\ifmath{{\mathrm{d}}} y_3}
\frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_1} - \frac{1}{\sigma^2_{{\mathrm{inel}}}} \frac{\ifmath{{\mathrm{d}}}^2\sigma}{\ifmath{{\mathrm{d}}} y_1\ifmath{{\mathrm{d}}} y_3}
\frac{\ifmath{{\mathrm{d}}}\sigma}{\ifmath{{\mathrm{d}}} y_2}
\end{eqnarray*}
\begin{eqnarray*}
{\mbox{with}} \ \ \ \ \ \sigma_{{\mathrm{inel}}} = \sum_{n\geq8}\sigma_n\ \ .
\end{eqnarray*}
The $\tilde C_{\ifmath{{\mathrm{S}}}}(y_1,y_2,y_3)$ correlation function is determined as a sum
of topological correlation functions:
\begin{equation}
\tilde C_{\ifmath{{\mathrm{S}}}} (y_1,y_2,y_3) = \sum_{n\geq8} P_n \tilde C_3^{(n)}
(y_1,y_2,y_3)\ ,
\end{equation}
\begin{equation}
\tilde C_3^{(n)} (y_1,y_2,y_3) = \tilde\rho_3^{(n)} (y_1,y_2,y_3) -
\tilde A_3^{(n)} (y_1,y_2,y_3)
\ , \end{equation}
\begin{eqnarray*}
\tilde A_3^{(n)} (y_1,y_2,y_3) = &\tilde\rho_2^{(n)}(y_1,y_2) \tilde\rho_1^{(n)}
(y_3)
+ \tilde\rho_2^{(n)}(y_2,y_3) \tilde\rho_1^{(n)}(y_1) + \tilde\rho_2^{(n)}(y_1,y_3)
\tilde\rho_1^{(n)}(y_2)\ - \\
& - 2 \tilde\rho_1^{(n)}(y_1) \tilde\rho_1^{(n)}(y_2) \tilde\rho_1^{(n)}(y_3)\ ,
\end{eqnarray*}
\begin{equation}
\tilde\rho_3^{(n)} (y_1,y_2,y_3) = \frac{1}{n(1,2,3)} \frac{1}{\sigma_n}
\frac{\ifmath{{\mathrm{d}}}^3\sigma}{\ifmath{{\mathrm{d}}} y_1\ifmath{{\mathrm{d}}} y_2\ifmath{{\mathrm{d}}} y_3}\ \ ,\end{equation}
\noindent
where $n(1,2,3)$ is the mean number of three-particle combinations in
events with charge multiplicity $n$.
The corresponding normalized function is defined as:
\begin{equation}
\tilde K_{\ifmath{{\mathrm{S}}}}(y_1,y_2,y_3) = \tilde C_{\ifmath{{\mathrm{S}}}}(y_1,y_2,y_3)/\sum_n
P_n\tilde\rho_1^{(n)}(y_1) \tilde\rho_1^{(n)}(y_2)\tilde\rho_1^{(n)}(y_3)\ .\end{equation}
Because of small statistics, three-particle correlations were not
observed in pp interactions at 200 GeV/$c$ at FNAL~\cite{Whit76}.
In $\PK^-\Pp$ interactions at 32 GeV/$c$~\cite{Buma79},
three-particle correlations were considered
using $\tilde C_{\ifmath{{\mathrm{S}}}}(y_1,y_2,y_3)$ and $\tilde K_{\ifmath{{\mathrm{S}}}}(y_1,y_2,y_3)$.
No positive short-range correlation effect was observed. Correlations
in the form of $K$ have been observed in the central region by the ISR
experiment for $n\geq8$~\cite{BreakEP88} .
Fig.~3.23 from NA22 shows $K_3(0,0,y)$ and $\tilde K_{\ifmath{{\mathrm{S}}}}(0,0,y)$ for
the combined M$^+$p sample at 250 GeV/$c$~\cite{Aiva91}. Also shown are the
values of $K_3(0,0,y)$ obtained in pp-interactions at $\sqrt s$=31-62 GeV
{}~\cite{BreakEP88} (lines). Inclusive three-particle correlations $K_3(0,0,y)$
are indeed seen in the NA22 data.
They are strongest when a third particle partially compensates the charge of
a pair of identical particles. There are, however, no correlation effects
visible in
the function $\tilde K_{\ifmath{{\mathrm{S}}}}(0,0,y)$. In FRITIOF and QGSM, three-particle
rapidity correlations are absent
in both $K_3(0,0,y)$ and $\tilde K_{\ifmath{{\mathrm{S}}}}(0,0,y)$.
Recently, a factorization of the normalized
three-particle correlation function has been proposed
{}~\cite{CaSa89,CapFiaKrz89,WolfAPP90} under the form of
a ``linked-pair'' structure:
\begin{equation}
K_3(y_1,y_2,y_3)=K_2(y_1,y_2)K_2(y_2,y_3)+K_2(y_1,y_3)K_2(y_3,y_2).
\label{linked:pair}
\end{equation}
The comparison of the prediction of (\ref{linked:pair}) to the data is given
in Table 3.2, for $n\geq 2$, at a resolution of 0.5 rapidity units.
At this resolution, the linked-pair ansatz is in agreement with the
measured three-particle correlation within two standard deviations.
Note, that $y$-correlations are much stronger for low-$p_\ifmath{{\mathrm{T}}}$ particles
and that the linked-pair ansatz continues to hold.
\vskip 2mm
With the accuracy presently attainable for three-particle correlations, it
is obvious that studies of still higher-order correlation functions require
better methods. The most successful ones will be discussed in Chapter~4.
\vspace{5mm}
{\bf Table 3.2}
The 3-particle correlation function compared to the
prediction from the linked-pair ansatz,
for non-single diffractive data ($n\geq 2$).
\vspace{5mm}
\begin{center}
\begin{tabular}{lllll}
\hline
&\multicolumn{2}{c}{all $p_\ifmath{{\mathrm{T}}}$} & \multicolumn{2}{c}{$p_\ifmath{{\mathrm{T}}}<$0.15
GeV/$c$}\\
&~~~~data &~~~~LPA &~~~data &~~~LPA \\
\hline
$K_3^{---}(0,0,0)$ & 0.23$\pm$0.10 &0.30$\pm$0.03 &2.3$\pm$1.7
&2.0$\pm$0.4 \\
$K_3^{+++}(0,0,0)$ & 0.14$\pm$0.06 &0.21$\pm$0.02 &1.2$\pm$0.6
&1.0$\pm$0.2 \\
$K_3^{\ifmath{{\mathrm{c}}}\rc\ifmath{{\mathrm{c}}}}(0,0,0)$ & 0.39$\pm$0.04 &0.53$\pm$0.03 &1.9$\pm$0.5
&1.7$\pm$0.2 \\
\hline
\end{tabular}
\end{center}
\section{Summary and conclusions}
\begin{enumerate}
\item The main contributions to the correlation functions $C_2$ and $C_3$
come from the mixing of events with different multiplicity and different
single-particle density, but some effect remains in the so-called short-range
correlation part.
\item $C_2(0,y_2)$ increases much faster with increasing energy than its
short-range contribution.
\item The short-range correlation is significantly larger for $(+-)$ than
for the equal-charge combinations, and is positive over a wider rapidity
range in
$C_2(y_1, y_2=y_1)$.
\item The correlation functions
$\tilde C_2^{(n)}(0,y_2)$,
contrary to $C^{(n)}_2(0,y_2)$,
are similar for different multiplicity $n$,
except that $\tilde C_2^{(n)+-}$ becomes narrower with increasing $n$.
\item In hadron-hadron collisions,
the correlation functions depend strongly on transverse momentum
and are largest for small-$p_\ifmath{{\mathrm{T}}}$ particles. Consequently, correlations are
stronger in multi-dimensional phase space than in a lower dimensional
projection,
such as rapidity space. Further implications of this observation will be
discussed in Sect.~4.3.
\item In the central c.m.~region, and at comparable energy,
the correlation strength observed in M$^+$p collisions at
$\sqrt s$=22 GeV is of similar magnitude as in $\re^+\re^-$ collisions
and as in $\mu$p collisions, if the trend of the latter is extrapolated
to $W=22$ GeV.
Model predictions for $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ and $\mu$p interactions
slightly underestimate the correlation strength but give, nonetheless,
clear evidence that (hard) gluon effects are the main source
of correlations in rapidity space.
\item Combinatorial background can be suppressed by studying the
correlation of strange particles. Data are scarce,
but support the conclusions drawn from data on non-strange particles.
\item The UA5 cluster Monte Carlo and FRITIOF describe
$C_2^{(n)}(\eta_1,\eta_2)$ at CERN-Collider energies, at least in the
charge multiplicity range $34\leq n\leq 38$.
At lower energies $(20\ ^<\hs-2.5mm_\sim\ \sqrt s \ ^<\hs-2.5mm_\sim\ 30$ GeV), the single-chain
LUND model shows a strong anti-correlation among like-charge particles.
The two-string FRITIOF model
and DPM predict negative values for $C_2(y_1,y_2)$ or $K_2(y_1,y_2)$
in the central region for like charges.
They are positive but far below the data
for unlike-charge pairs.
QGSM reproduces $C_2(y_1,y_2)$ and $\tilde C_2(y_1,y_2)$
for all charge combinations, but cannot account for the short-range part
$\tilde C_\ifmath{{\mathrm{S}}}(y_1,y_2)$.
\item Positive correlations are observed at large values of the azimuthal
angle $\Delta\varphi$, as expected from transverse momentum conservation.
The correlations among like-charge particle pairs at small values of
$\Delta\varphi$ and $\Delta y$, where Bose-Einstein effects should contribute, are
significantly larger than predicted by FRITIOF~2, DPM and QGSM.
The deviations are stronger for particles with small transverse momentum.
\item In general, short-range correlations in $\re^+\re^-$ annihilation are
reasonably well described by the LUND- and Webber-type models. To the
contrary, in models for hh collisions which contain only one or two strings
without additional $p_\ifmath{{\mathrm{T}}}$ effects or gluons, correlations in rapidity
are underestimated, those in azimuthal angle overestimated.
Models such as LUND and DPM are known to underestimate the height
of the ``sea-gull'' wings (the particle average transverse momentum as
a function of Feynman-$x$)~\cite{Ajin87}, a signal of semi-hard
interactions. The models neglect such processes.
This may partially explain why the models fail in both instances.
\item The distribution in the interparticle opening angle of $\re^+\re^-$ collisions
at LEP favours models with coherent parton showering.
\item Three-particle correlations are now observed in all charge combinations.
They are particularly large for low-$p_\ifmath{{\mathrm{T}}}$ particles. Within two standard
deviations, they satisfy the linked-pair ansatz. No short-range contribution
$\tilde K_{\ifmath{{\mathrm{S}}}}$ is observed in three-particle correlations.
Other methods are needed to study higher-order correlations.
\end{enumerate}
\chapter{Multiplicity fluctuations and intermittency}
\section{Prelude}
The study of fluctuations in particle physics already has a
long history going back to early
cosmic-ray observations. To our knowledge, Ludlam and Slansky~\cite{Ludlam73}
were the first to advocate analysis of event-to-event
fluctuations in hadron-hadron collisions.
Comparing rapidity distributions of single events with the
sample averaged distribution, they put in evidence
strong clustering effects in longitudinal phase-space, indicating
``a remarkably structured phase-space density''~\cite{Slansky74}.
Fluctuations in individual events were also considered in the
context of Reggeon theory in the important paper establishing the
AGK-cutting rules~\cite{Abramov73}.
Early evidence for large concentrations of the particle number in small
rapidity regions for single events were reported in
cosmic ray experiments~\cite{Ale62,Ara78,Sla81} and
in pN collisions at 200 GeV beam momentum~\cite{Maru79}.
A further number of high density ``spikes" in
rapidity space have been reported during the last decade.
Fig.~4.1a shows the notorious JACEE event~\cite{Burn83} at a pseudo-rapidity
resolution (binning) of $\delta\eta=0.1$. It has
local fluctuations up to $dn/\delta\eta\approx300$ with a signal-to-background
ratio of about 1:1.
The NA22 event~\cite{AdamPL185-87} of Fig.~4.1b
contains a ``spike" at a rapidity resolution $\delta y=0.1$
of $dn/dy=100$, corresponding to 60 times the average
density in this experiment. UA5~\cite{Carl87} has
reported ``spikes" in $dn/d\eta$ up to 30 (10 times average) as early as
JACEE, but found these to be in agreement with a short-range cluster
Monte Carlo. Also EMU-01~\cite{Adamo88} sees events with
$dn/d\eta=140$ satisfactorily explained by FRITIOF.
{}From an experimental point of view, there is little doubt that events
with large local density fluctuations exist. The real
question is whether these are of dynamical or merely statistical origin,
whether the underlying probability density is continuous or intermittent.
Early attempts were made to answer the
question of non-statistical fluctuations employing
transform techniques~\cite{Takagi}, but these were not followed up.
The problem resurfaced
in the work of {Bia\l as} and Peschanski~\cite{bialas1,bialas2}, who
suggested that spikes could be a manifestation in hadron physics
of ``intermittency'', a phenomenon well-known in fluid-dynamics.
The authors argued that if intermittency occurs in particle production,
large density fluctuations are not only expected, but should also exhibit
self-similarity with respect to the size of the phase-space volume.
Ideas on self-similarity and fractals in jet physics had earlier been
formulated in~\cite{Fey79,Ven79}, rephrased in the
language of QCD branching processes in~\cite{Kon79} and in a
simplified form in~\cite{Gio79}.
For soft hadronic processes, fractals and self-similarity were
first considered in~\cite{Minh83} and their quantitative measures
in~\cite{DremJETP87,DremFest}.
In multiparticle experiments, the number of hadrons produced
in a single collision is small and subject to considerable
``noise''. To exploit the techniques employed in complex system
theory, a method must be devised to separate fluctuations of purely
statistical origin, due to finite particle numbers, from
the possibly self-similar fluctuations of the underlying particle
densities. The latter are the quantities of physical interest.
A solution, already used in optics and suggested for multiparticle
production in~\cite{bialas1,bialas2},
consists in measuring suitably normalized factorial moments
of the multiplicity distribution in a given phase-space volume.
\section{Normalized factorial moments}
\subsection{The method}
The method proposed in~\cite{bialas1,bialas2}
consists in measuring the dependence of the normalized
factorial moments $F_q(\delta y)$ defined
in (\ref{dr:44}-\ref{dr:46}) as a function of the resolution $\delta y$.
For definiteness, $\delta y$ is supposed to be an interval in rapidity,
but the method generalizes to arbitrary phase-space dimensions.
In Sect.~2.2 we have pointed out that the scaled factorial moments
enjoy the property of ``noise-suppression".
It is easily verified that this crucial property does not apply
to ordinary moments $\langle n^q\rangle/\langle n\rangle^q$ (cfr.~Sect~4.7 below).
High-order moments further act as a filter and
resolve the large $n_m$ tail of the multiplicity distribution.
They are thus particularly sensitive to large density fluctuations
at the various scales $\delta y$ used in the analysis.
As proven in~\cite{bialas1,bialas2}, a ``smooth" (rapidity) distribution,
which does not show any fluctuations except for the statistical ones,
has the property that $F_q(\delta y)$ is independent of the resolution $\delta y$
in the limit $\delta y\to 0$. This follows directly from (\ref{dr:55}), if
$P_\rho$ is a product of $\delta$-functions in
$\rho_m\ (m=1,\dots,M)$ centered around $\langle\rho_m\rangle$.
On the other hand, if dynamical fluctuations exist and $P_\rho$ is
``intermittent'' (i.e., regions of fluctuations exist at all scales of $y$),
the $F_q$ obey the power law (\ref{fq113}).
Equation (\ref{fq113}) is a scaling
law, since the ratio of the factorial moments at resolutions $L$ and $\ell$
\begin{equation}
R = F_q(\ell)/F_q(L) = (L/\ell)^{\phi_q}
\end{equation}
only depends on $L/\ell$.
As mentioned in Sect.~2.4 and Subsect.~5.2.2, the ``intermittency indices''
$\phi_q$ (slopes in a double-log plot) are related~\cite{Hen83,LiBu89,Hwa90}
to the anomalous dimensions $d_q=\phi_q/(q-1)$, a measure for the
deviation from an integer dimension.
We noted in Sect.~3.4 that the experimental study of correlations
is difficult already for three particles. The close connection between
correlations and factorial moments (Subsect.~2.1.4) offers a possibility to
measure higher-order correlations with the factorial moment
method at smaller distances than previously feasible.
The method further relates possible scaling behaviour
of such correlations to the physics of fractal objects.
Despite the advantages, it should be remembered that reliable data
can only be extracted if factorial moments are averaged over a
large domain of phase space. This holds the
danger of obscuring important dynamical effects.
The definition of ``intermittency" given in (\ref{fq113}), has its
origin in other disciplines\footnote{\ For
a masterly expos\'e of this subject see~\cite{Zeld90}}.
It rests on a loose parallel
between the high non-uniformity of the distribution
of energy dissipation, for example, in turbulent intermittency and the
occurrence of large ``spikes" in hadronic multiparticle final states
(Sect.~4.1). In the following we use
the term ``intermittency'' in a weaker sense, referring to
the rise of factorial moments with increasing resolution but not necessarily
according to a strict power law.
The suggestion that normalized factorial moments of particle
distributions might show power-law behaviour has spurred a vigorous
experimental search for (more or less) linear dependence of $\ln F_q$
on $-\ln\delta y$. Within a surprisingly short time (one-dimensional) analyses
were performed for $\re^+\re^-$
{}~\cite{BuLiPe88Aba90,Brau89,Behr90,Abreu90,Akr91,Dec91,Mur93},
$\mu$p~\cite{Dera90},
$\nu A$~\cite{Verlu90},
hh~\cite{Ajin89-90,Alba90,Singh91,Are91,Bravi,Rimon91,Wang94},
h$A$~\cite{Holy89,Dera90Sing,Bott91,Ghosh92,Shiv93,Shiv94} and
$AA$~
\cite{Holy89,Dera90Sing,Adamo90,Seng90,AAke90,Ghosh91,Sark93,Abbo94,Albr94}
collisions.
With respect to the original objective, the early one-dimensional work
has remained inconclusive, but valuable information and experience was
accumulated. Much more promising insight has come from studies in two-
and three-dimensional phase space. This is discussed in Sect.~4.3.
Further extensions of this approach, concentrating on improved integration
methods and differential studies in Lorentz-invariant variables have lead
to further clarification of the issues involved in intermittency.
These very recent developments are presented in Sects.~4.8-4.10.
\subsection{Results on log-log plots (in one dimension)}
In this and the next few sections we review experimental results
and model predictions obtained from one-dimensional studies.
Due to the vast amount
of data available, we limit ourselves to an illustration of the
major characteristics of factorial moment behaviour in
various processes and at various energies.
In Fig.~4.2a, log$F_5$ is plotted~\cite{bialas1,bialas2}
as a function of -log $\delta\eta$ ($\eta$ is the pseudorapidity)
for the JACEE event. It is compared with an independent
emission Monte-Carlo model tuned to reproduce the average $\eta$ distribution
of Fig.~4.1a and the global multiplicity distribution, but has no short-range
correlations included. While the Monte-Carlo
model indeed predicts constant $F_5$, the JACEE event shows a first
indication for a linear increase, i.e. a possible sign of intermittency.
Further examples are given in Fig.~4.2b for KLM~\cite{Holy89}, again showing
an roughly linear increase for $\delta\eta<1\ (-\ln\delta\eta>0)$ instead of
the flat behaviour expected for independent emission, and in Figs.~4.2c and d
for UA1 \cite{Alba90} in terms of $\delta\eta$ and $\delta\phi$, respectively.
Anomalous dimensions $d_q$ fitted over the range $0.1<\delta y(\delta\eta)<1.0$
are compiled in Fig.~4.3~\cite{Bial90}. They typically range from
$0.01$ to $0.1$, which means that the fractal (R\'enyi) dimensions $D_q=1-d_q$
are close to one. The $d_q$ are larger and grow faster with increasing order
$q$ in $\mu$p and $\re^+\re^-$ (Fig.~4.3a) than in hh collisions (Fig.~4.3b)
and are small and almost independent of $q$ in heavy-ion collisions
(Fig.~4.3c). For hh collisions, the $q$-dependence is considerably stronger
for NA22 ($\sqrt{s}=22$ GeV) than for UA1 ($\sqrt{s}=630$ GeV).
\subsection{Model predictions}
\subsubsection{Hadron-hadron collisions}
A comparison to NA22 data on slopes $\phi_q$ (Fig.~4.4a) shows~\cite{Ajin89-90}
that intermittency is absent at $\sqrt{s}=22$~GeV in a two-chain DPM and
underestimated by FRITIOF. In Fig.~4.4b, PYTHIA is seen to stay below the
UA1 data~\cite{Alba90}, even after inclusion of Bose-Einstein interference
for identical particles. The UA5 cluster Monte Carlo GENCL, able to reproduce
conventional short-range correlations (at least in a certain range of
multiplicities cfr.~Fig.~3.3), follows the data down to a resolution
of $\delta\eta\approx0.3$, but completely fails for smaller $\delta\eta$.
Also, a multi-chain version of DPM including mini-jet production has
been compared to NA22 and UA1 data. The slopes are found to be too small
by at least a factor of~2~\cite{Bopp90}.
With respect to intermittency analysis, the situation may improve
with the introduction of ECCO~\cite{Hwa92}, an eikonal cascade model
based on geometrical branching, which now can account for strong
fluctuations, in particular in higher dimensions (Sect.~4.3 below).
However, the present version of ECCO is still less refined
than the more conventional models with respect to other observables.
The above examples show that present models for multiparticle production
in hh collisions are unable to reproduce the magnitude and the growth
of factorial moments with increasing resolution. From the discussion in
Chapter~3, it is evident that model predictions for correlations in general
are quite unreliable. The two-particle correlation function, measured by
$F_2$, also determines to a large extent the higher-order factorial moments
(cfr. Eq.~\ref{dr:48}) because of the weakness of genuine high-order
correlations. It is, therefore, mandatory to improve the models before
evidence for ``new physics" at very small (rapidity) separation can be
claimed. We return to this important question in later sections.
\subsubsection{h$A$ and $AA$ collisions}
The intermittency indices are much smaller in h$A$ and $AA$ collisions
than in hh collisions, and the event samples are much smaller. Model
comparisons are, therefore, less conclusive than in hh collisions. FRITIOF
is found too low in NA22 \cite{Bott91} for $\pi^+/\PK^+$ on Al and Au at
250 GeV/$c$, in E802 \cite{Abbo94} for central $^{16}$OAl and $^{16}$OCu at
14.6 $A$ GeV/$c$, in WA80 \cite{Albr94} for SS and Au at 200 $A$ GeV/$c$,
and in NA35 \cite{Dera90Sing} for pAu, OAu, SAu and SS at 200 $A$ GeV/$c$.
In WA80 it is shown that rough agreement can be obtained by renormalization
to the leftmost point of FRITIOF on the log-log plot (essentially the shape
of the overall multiplicity distribution) to the data. NA35 shows that
agreement can be obtained by adding Bose-Einstein interference for
like-charged particles (for a detailed analysis of the influence of BE
correlations see further below).
\subsubsection{Lepton-hadron collisions}
In Fig.~4.5a EMC data~\cite{Dera90} are compared to what is expected from
an extrapolation of conventional short- and long-range
correlations~\cite{CapFiaKrz89}. At small $\delta y$, the data are consistently
above these expectations. As Fig.~4.5b shows, the slopes $\phi_q$ in the same
data are considerably larger than predicted by the Webber and LUND models.
Similarly, Fig.~4.5c shows too low $\ln F_3$ from LUND, not only for
$\nu$Ne but also for the ``simpler'' $\nu \ifmath{{\mathrm{D}}}_2$ interactions~\cite{Verlu90}.
We tentatively conclude that also presently used {\it lepton-hadron} models
as such are unable to reproduce the intermittency observed in this process.
\subsubsection{$\re^+\re^-$ annihilation}
The annihilation of $\re^+\re^-$ into hadrons is by far the best understood of all
multihadron reactions. Creation of hadrons is traditionally pictured as a
multi-step process comprising a ``hard" parton evolution phase, described by
perturbative QCD---the parton shower---and a non-perturbative colour-confining
soft hadronization phase (Fig.~4.6). The former is a cascade process of nearly
self-similar type, and is expected to show characteristics typical of a
fractal object~\cite{Fey79,Ven79,Gio79}. In fact, already in 1979, in a
discussion of QCD jets, it was stated~\cite{Ven79} that ``the resulting
picture of a jet is formally similar to that of certain mathematical objects,
known as fractals, which look more and more irregular and complex as we
look at them with a better and better resolution". The expectation is,
therefore, that parton showers should exhibit intermittency at the parton
level. However, this is not sufficient to guarantee ``intermittency'' at
the hadron level. It is indeed difficult to imagine how the ``re-shuffling''
of the parton momenta during the hadronization phase with, e.g., the
formation of hadronic resonances and their subsequent decay would preserve
the (supposedly singular) nature of the correlations. A local parton-hadron
duality type of explanation is not satisfactory either, since ``it is merely
a name for a mechanism that is not at all understood''~\cite{Bial92}.
To describe the hadronization phase, all present Monte-Carlo codes rely
in last instance on a large amount of $\re^+\re^-$ data at different energies and
are carefully tuned to these. It came, therefore, as a surprise that a first
(indirect) analysis~\cite{BuLiPe88Aba90} of HRS results, shortly followed by
TASSO data~\cite{Brau89}, revealed deviations from model predictions quite
similar to those observed in lh and hh collisions (Figs.~4.7a,b).
More recently, CELLO~\cite{Behr90} and, in particular, the LEP
experiments~\cite{Abreu90,Akr91,Dec91}, claim ``reasonable''
agreement with the parton shower version of the LUND Monte Carlo
(Figs.~4.7c,d). Nevertheless, new DELPHI data now show, with ten times
larger statistics, significant deviations even
with a ``re-tuned" version of the Monte Carlo (Fig.~4.7d).
The origin of intermittency in the models is not quite as clear as is often
stated. Indeed, comparison of the factorial moments on parton and hadron
level in Figs.~4.8a,b~\cite{BotBusch}, shows that in (standard) JETSET the
increase of $\ln F_q$ at small $\delta y$ is not due to the parton shower, but to
hadronization! Only if the parton shower is allowed to continue down to very
low $Q^2_0$ values (Fig.~4.8c,d for $Q^2_0$=0.4 GeV$^2$), implying local
hadron-parton duality, is intermittency becoming visible also at the parton
level. It has been verified that the influence of $Q^2_0$ is, of course,
much less important at $1$~TeV.
On the other hand, intermittency seems to be fully developed on the parton
level already at 91 GeV in the Webber model, and is in fact smeared out by
hadronization~\cite{Jedr89}.
The sensitivity to the cut-off in the perturbative QCD cascade and the
role of hard and soft phases has also been discussed
in terms of the dipole radiation model~\cite{Dahl89}.
Intermittency can be increased in the soft phase by an increase of the
$\pi/\rho$ ratio, also required from direct measurements by NA22~\cite{Agab90},
EMC~\cite{ArneC33-86} and in hA collisions~\cite{Walker91}. The direct pions
resolve the underlying parton structure better than the more massive
resonances. From a tunnelling production mechanism, these pions are expected
to have smaller $p_\ifmath{{\mathrm{T}}}$ than other particles, a property presently neglected
in the MC programs. A Goldstone-like mechanism causing additional soft
direct pion production at break-up points has recently been suggested by
the LUND group~\cite{AndGu94}.
For further progress, additional studies are needed.
$\bullet$ One should identify the true causes of intermittency in present
Monte-Carlo models, preferably on more sensitive distributions, such as those
to be discussed below.
This should reveal the influence of hard and soft gluon
emission at high energies where parton showering is fully developed and
dominates over the soft phase.
$\bullet$ Intermittency is also particularly sensitive to the exact
treatment of the {\it soft} phase.
This phase can be studied with high statistics
at lower energies where parton showering is less important.
\subsection{A warning}
Before going into the necessary further detail,
we should mention the influence
of possible experimental biases. On purpose and by its very definition,
the higher factorial
moments are sensitive to a small number of events in the tail of the
multiplicity
distribution in small phase-space bins.
Moments can be {\it reduced} by limited two-track resolution, by track losses
from limited acceptance or bad reconstruction, or simply due to truncation of
the multiplicity distribution in a finite event sample.
Moments can be {\it increased} due to double counting of tracks (track match
failures), Dalitz decays and nearby $\gamma$-conversions or K$^0/\Lambda$ decays. A
dangerous increase comes from the commonly used ``horizontal'' averaging,
where a {\it constant} average (pseudo-) rapidity distribution is assumed
over the range $\Delta Y$. Contrary to first belief, this problem is {\it not}
completely solved by the correction method proposed in~\cite{Fial89} !
Further influence is to be expected from the choice of the sample, e.g.,
inelastic or non-diffractive, cuts on multiplicity, cuts on $p_\ifmath{{\mathrm{T}}}$,
all events or only those with $n \geq n_0$ in $\Delta Y$, etc., the size and
position of $\Delta Y$, the $\delta y$ region chosen for the fit and the correlation
of errors.
Many of these effects have been studied in a number of experiments and we
refer to these and to~\cite{Wo90Ek90Sei90,Frie89} for more details.
\section{Higher dimensions}
\subsection{The projection effect}
So far, we have discussed factorial moments derived from one-dimensional
distributions in rapidity or pseudorapidity. The analysis can evidently be
extended to other 1D variables, such as the azimuthal angle $\varphi$ in the plane
perpendicular to the beam or event axis, or the particle transverse momentum
($p_\ifmath{{\mathrm{T}}}$). Given sufficient statistics, distributions can be analysed in two-
and three-dimensional phase-space domains. Common choices are $(\Delta y,\Delta\varphi)$,
$(\Delta y,\Delta\ln p_\ifmath{{\mathrm{T}}})$, $(\Delta\varphi,\Delta\ln p_\ifmath{{\mathrm{T}}})$ and $(\Delta y,\Delta\varphi,\Delta\ln p_\ifmath{{\mathrm{T}}})$.
Fig.~4.9a gives an example of 1D-results from UA1~\cite{Alba90} showing that
intermittency is also present in $\varphi$. The intermittency effect is larger
when two-dimensional cells $(\Delta y,\Delta\varphi)$ are studied than in 1D
(Fig.~4.9b,c). This is particularly pronounced in $\re^+\re^-$ annihilations
(Figs.~4.10a,b), the measured slopes $\phi_q$ being about six times larger
in 2D than in 1D. These observations
are now understood to imply that intermittency ``lives in
3D"~\cite{Ochs,BiaSei90}. Projection onto lower-dimensional subspaces dilutes
the effect and leads to flattening of the factorial moments. This is most
pleasantly demonstrated by the fact that one can enjoy a continuous
(two-dimensional) shadow of a tree, in spite of the self-similar branching of
this tree in three dimensions.
The projection-effect is convincingly illustrated in Fig.~4.10a,b. The lines
in sub-fig. a) are fits by a 2D $\alpha$-model; the curves in sub-fig. b) are the
projections onto rapidity-space and show considerably less increase and even
a flattening for $\delta y\to 0$ (note the difference in scale). Nevertheless,
Fig.~4.10a still shows saturation of $F_2$ at large $M$ even in 2D, an
indication that an analysis in three dimensions may be required.
\subsection{Transformed momentum space}
To study intermittency in three-dimensional phase space, one faces the
additional difficulty that the particle density is all but uniform in the
usual single-particle variables $y,\varphi$ and $p_\ifmath{{\mathrm{T}}}$. The distribution in
$p_\ifmath{{\mathrm{T}}}$ is in fact falling exponentially. Uniformity of the density is,
however, an explicit assumption in the derivation of the power law
(\ref{fq113}). Violation of this condition renders an intermittency analysis
useless.
To circumvent this problem, the authors of~\cite{Ochs,BiGa90} have proposed
to use domains in a transformed momentum space with (almost) constant
density.
This is accomplished by a transformation of the original variables $y,\varphi$
and $\ln p_\ifmath{{\mathrm{T}}}$ to ``cumulative" variables. Thus, for a single variable,
say $y$, one defines the new variable $X(y)$ as
\begin{equation}
X(y) = \frac{\int^y_{y_{{\mathrm{min}}}}\rho_1(y')dy'}{\int^{y_{{\mathrm{max}}}}_{y_{{\mathrm{min}}}}
\rho_1(y')dy'}\ \ .
\end{equation}
For higher dimensions, it is assumed in~\cite{Ochs} that the single particle
density factorizes as
\begin{equation}
\rho_1(y,\varphi,p_\ifmath{{\mathrm{T}}}) = \rho_\ifmath{{\mathrm{a}}}(y)\rho_\ifmath{{\mathrm{b}}}(\varphi)\rho_\ifmath{{\mathrm{c}}}(p_\ifmath{{\mathrm{T}}})\ \ .\end{equation}
\noindent
Under this rather strong hypothesis, one can transform each of the three
variables independently. The method proposed in~\cite{BiGa90} does not assume
factorization but is technically quite involved. In practice, the two
techniques give satisfactorily similar results~\cite{Bott92}.
Data on $F_2$ in various dimensions are shown in Fig.~4.11 for
$\re^+\re^-$~\cite{Abreu90} and hh collisions~\cite{Ajin89-90,Alba90}. In all cases,
the data behave more power-like in 2D than in 1D. From Fig.~4.11a, it is also
evident that JETSET PS remains in good agreement with $\re^+\re^-$ data in higher
dimensions.
At variance with power-law behaviour expected from intermittency, NA22 finds
that the 3D factorial moments show an upward bending (Fig.~4.11c). This
effect
persists after exclusion of Dalitz decays and $\gamma$-conversions. A rise faster
than power law is also observed in 3D for collisions of various projectiles
with $\mbox{Au}$ by NA35 (Fig.~4.12a)~\cite{Dera90Sing}. Following a
suggestion
in~\cite{FiMPI91}, NA35 finds that the normalized factorial cumulant
$K_2=F_2-1$ shows much better linearity in a log-log plot than $F_2$ itself
(Fig.~4.12b).
This observation, in fact, furthers considerably our understanding of the
intermittency phenomenon. In~\cite{FiMPI91,Fial91} the author has compared
3D data on $F_2$ at $\sqrt s\simeq20$ GeV for $\mu$p~\cite{Dera90},
$\pi/\PK$p~\cite{Ajin89-90}, $\mbox{pAu}$, $\mbox{OAu}$ and
$\mbox{SAu}$~\cite{Dera90Sing} collisions using the parametrization
\begin{equation}
F_2 = 1 + c (M^3)^{\phi_2} + c' \ \ ,
\label{eq:3.15}
\end{equation}
where $M^3$ is the number of 3D phase-space cells. The second term in
(\ref{eq:3.15}) is equal to $K_2$. The constant $c'$ accounts for
long-range correlations, known to exist in hh collisions.
The comparison of (\ref{eq:3.15}) to the data is shown in Fig.~4.13; the
parameters are given in the figure caption. The parameter $c'$ is negligible
for $\mu$p and heavy ion collisions, but non-zero for meson-proton and p$A$
collisions, in agreement with expectations. The most noteworthy result,
however, concerns $\phi_2$, which is seen to have a value in the range 0.4-0.5
for all processes. This is remarkable in various respects.
Firstly, if confirmed in further studies, and in particular for $\re^+\re^-$
annihilation, it strongly suggests that the resolution-dependence of $F_2$
exhibits a high degree of ``universality", is independent of specific
details of the production process and thus reflects general features of
hadronization dynamics.
Secondly, such universality is at variance with the hitherto accepted idea
that the factorial moments and the anomalous dimensions become the smaller
the more complex the collision process, due to an increasing inter-mixing of
production sources~\cite{LiBu89}.
Thirdly, if ``universality" continues to hold in high energy $\re^+\re^-$
annihilation,
one must revise the commonly expressed opinion that the perturbative parton
evolution, and in particular hard-jet emission, is the primary cause of the
rise of factorial moments at high resolution. Needless to say, it would be
most interesting to verify systematically the universality conjecture in
other reactions and for three-particle correlations.
The experimental success of expression (\ref{eq:3.15}) becomes quite
intriguing when one realizes that the volume $\delta\sim M^{-3}$ of a
phase-space cell (for sufficiently large $M$) is in fact related to
the invariant mass $M_{{\mathrm{inv}}}$ of the two-particle system or to $Q^2$,
the square of their four-momentum difference. The form
(\ref{eq:3.15}) implies that the two-particle correlation function
behaves as a power law in $M_{{\mathrm{inv}}}$ or $Q^2$.
The data, therefore, seem to tell that an intermittency analysis
should be performed in (Lorentz-invariant) multiparticle variables,
rather than single-particle variables. This will be further discussed
in Sect.~4.8-4.10.
As mentioned in Subsect.~4.2.3 above, ECCO~\cite{Hwa92} has some success
in describing the NA22 data on fluctuations in varying scales of
resolution, in particular when the analysis is done in three
dimensions (Fig.~4.14). The basis of this model is geometrical branching
for soft production at low $p_\ifmath{{\mathrm{T}}}$. The geometrical aspect of hadrons,
i.e. the fact that they are extended objects, puts the impact
parameter $R$ in a pre-eminent role. The fluctuation in $R$ from event
to event leads to fluctuations in $p_\ifmath{{\mathrm{T}}}$ and explains the
non-vanishing intermittency in $\ln p_\ifmath{{\mathrm{T}}}$ reported by NA22. The
(stronger) intermittency in rapidity can be generated only with a
singular splitting function for branching in rapidity space. Since
there is no branching in $\varphi$ in the model, intermittency is nearly
non-existent in this variable. Still, the long-range correlation due
to $p_\ifmath{{\mathrm{T}}}$ conservation leads to a decrease of $F_q$ at large bin
size, a feature also observed by NA22.
\subsection{A generalized power law}
It has been pointed out~\cite{OchWo88-89} that
the one-dimensional moments follow the generalized power law
\begin{equation}
F_q \propto (g(\delta y))^{\phi_q}\ ,
\label{ochs:rel}
\end{equation}
in multiplicative cascade models.
In (\ref{ochs:rel}), $g(\delta y)$ is a
general function of $\delta y$. Expressing $g$ in terms of
$F_2$, one finds the linear relation
\begin{equation}
\ln F_q=c_q+(\phi_q/\phi_2)\ln F_2 \ ,
\label{ochs:rel1}
\end{equation}
from which the ratio of anomalous dimensions is directly obtained.
This intriguing relation has successfully been confirmed by
experiment, not only in one dimension, but up to 3D~\cite{Ochs}.
Moreover, the ratios $\phi_q/\phi_2$ are found to be largely independent
of the dimension of phase space~(Fig.~4.15a) and of the type of collision
(Fig.~4.15b).
The ratio of the anomalous dimensions $d_q(=\phi_q/(q-1))$
and $d_2$ are shown in Fig.~4.16b as a function of $q$.
The $q$-dependence is claimed to be
indicative of the mechanism causing intermittent behaviour.
For a (multiplicative) cascade mechanism,
in the log-normal approximation (long cascades), the moments satisfy the
relation~\cite{bialas1,bialas2}
\begin{equation}
\frac{d_q}{d_2}=\frac{\phi_q}{\phi_2} \frac{1}{q-1}=\frac{q}{2}.
\label{3:19}
\end{equation}
However, the use of the Central Limit Theorem for a multiplicative process,
such as in the $\alpha$-model, is a very crude approximation~\cite{AlbBi91}
particularly in the tails. As argued in~\cite{BrPe91}, a better description
might be obtained if the density probability distribution is assumed to be
a log-L\'evy-stable distribution, characterized by a L\'evy-index
$\mu$. In that case (\ref{3:19}) generalizes to
\begin{equation}
\frac{d_q}{d_2}=\frac{1}{2^\mu-2}\frac{q^\mu-q}{q-1}\ .
\label{3:19b}\
\end{equation}
For $\mu=2$, the Gaussian case, (\ref{3:19b}) reduces to (\ref{3:19}).
The multifractal behaviour characterized by (\ref{3:19}-\ref{3:19b})
reduces to a mono-fractal behaviour~\cite{Satz89,BialHwa91}
\begin{equation}
\frac{d_q}{d_2}=1
\end{equation}
for $\mu=0$, implying an order-independent anomalous dimension. This
would happen if intermittency were be due to
a second-order phase transition.
Consequently, monofractal behaviour
might be a signal for a quark-gluon-plasma phase transition.
The data are best fitted with the L\'evy-law solution with $\mu=1.6$. This
value is inconsistent with the Gaussian approximation,
and also definitely higher than expected
for a second-order phase transition.
The validity of the dimension-independent generalized power behaviour
has been questioned in a recent NA22 analysis~\cite{Ajin89-90} shown in
Fig.~4.16a. While a fit to the combined data on all variables and dimensions
(full circles), as well as a weighted average over all individual fits give
$\mu$ values in rough agreement with those of~\cite{Ochs}, the 3D-data have
$\mu>2$, not allowed in the sense of L\'evy-laws.
Even larger values of $\mu$, ranging from 3.2 to 3.5, have been found for
$\mu$p deep-inelastic scattering in~\cite{BrPe91}. According
to~\cite{Ratti91,Ratti92}, this is evidence that the procedure to obtain
the L\'evy-index is used outside its domain of validity. An allegedly more
general method, based on Double Trace Moments (to be discussed in
Subsect.~4.7.7) indeed yields $\mu$-values within the mathematically allowed
boundaries. However, we shall see that the latter method may be criticized
on other grounds. A possible way out is self-affinity to be discussed in
Subsect.~4.3.5, below.
The linear $d_q/d_2$ behaviour in Fig.~4.16a) and b) gives some justification
for (\ref{BB:134}). Fig.~4.16c) and d) show~\cite{chek94} the slope $r$
of (\ref{BB:134}) for a number of experiments. All experiments, except
perhaps SAg/Br, show multifractal behaviour $(r>0)$.
Despite the confusion, it remains a noteworthy experimental fact that
the factorial moments of different orders obey simple
hierarchical relations of the type~(\ref{ochs:rel1}).
This means that correlation functions of different orders are
not completely independent but are somehow interconnected.
Such situations are commonly encountered in various branches of
many-body physics (see e.g.~\cite{CaSa89,WolfAPP90,Peeb80}), but a
satisfactory link with particle phenomenology, let alone QCD, remains
to be established. Nevertheless, on a simple example it was recently
shown~\cite{edw:sant} that a linear relation between $\ln F_3$ and
$\ln F_2$ can be obtained if the connected correlation functions are
assumed to be of a factorized Mueller-Regge power-law form in two-particle
invariant-masses squared $s_{ij}$,
i.e.~ $C_3(1,2,3)\propto (s_{12})^{1-\alpha_1}\,
(s_{23})^{1-\alpha_2} +\,\mbox{cycl. perm.}$. Note that this Regge-form
has the ``linking'' structure of (3.13).
\subsection{Thermal versus non-thermal phase transition}
\subsubsection{Second order phase transition?}
A simple model that can provide some
hint on the nature of a second-order phase transition is the Ising model
in 2D \cite{Ising}. Its intermittency behaviour has been studied both
analytically and numerically \cite{Satz89,Wosiek88}. The anomalous dimension
is found to be $d_q=1/8$, independent of $q$. Based on that finding, it has
been conjectured that intermittency may be monofractal if due to a QCD
second-order phase transition \cite{BialHwa91}. However, as mentioned in
Subsect.~4.3.3 above, all types of interactions, including heavy-ion
collisions, show multifractal behaviour.
Of course, the Ising model is very simple and the above conjecture has little
basis. In \cite{hwanaz}, intermittency is, therefore, studied in the
framework of the Ginzburg-Landau theory also used to describe the confinement
of magnetic fields into fluxoids in a type II superconductor. In the model,
the anomalous dimension is not constant, but follows
\begin{equation}
\frac{d_q}{d_2} = (q-1)^{\nu-1}\ \ \ , \ \ \nu=1.304\ \ ,
\label{4.10}
\end{equation}
with $\nu$ being a universal quantity valid for all systems describable by the
GL theory, independent of the underlying dimension or the parameters of the
model. This is of particular importance for a QCD phase transition, since
neither the transition temperature nor the other important parameters are
known there.
In quantum optics, $\gamma$ production at the threshold of lasing is
discribable as a second-order phase transition. Indeed, a photo-count
experiment \cite{Young} has verified (\ref{4.10}) to high precision.
On the other hand, the current NA22 data on particle production in hadronic
collisions give $\nu=1.45\pm 0.04$ \cite{Charl}, heavy-ion experiments
$\nu=1.55\pm 0.12$ \cite{hwanaz} and $\nu=1.459\pm 0.021$ \cite{Seng90}.
For a first-order phase transition, all $d_q$ are zero and no intermittency
would be observed \cite{BialHwa91}.
However, it has been shown in \cite{Babi95} that in a generalized GL model,
a first-order phase transition combined with the quantum optics analogy
of lasing at threshold can lead to intermittency behaviour in some regions
of the parameters, with approximately the same intermittency indices as
a second-order phase transition.
\subsubsection{Non-thermal phase transition?}
Of course, the phase transition does not need to be thermal, i.e.,
the new phase need not be characterized by a thermodynamical behaviour.
Such a transition could, e.g. take place during a parton-shower cascade
and has been formulated in \cite{VHov89} for a number of ``ultra-soft''
phenomena, including intermittency. It leads to the co-existence of different
phases, in analogy to different phases of the spin-glass systems. The
examples
of the JACEE event (Fig.~4.1a), which contains many ``spikes'' and
``holes'', and that of the NA22 event (Fig.~4.1b), which consists of just
one spike, indicate that such a possibility may be more than just a
speculation.
The condition for the existence of such different phases of a self-similar
cascade is that the function
\begin{equation}
\lambda_q = (\phi_q+1)/q
\label{4.11}
\end{equation}
has a minimum at some value $q=q_\ifmath{{\mathrm{c}}}$ (not necessarily an integer)
\cite{PescTH5891,Pesc89,BrPe90,BiaZa}. The regions $q<q_\ifmath{{\mathrm{c}}}$ and $q>q_\ifmath{{\mathrm{c}}}$
are dominated by numerous small fluctuations and rare large fluctuations,
respectively. In the terminology of \cite{BiaZa}, the system resembles a
mixture of a ``liquid'' of many
small fluctuations and a ``dust'' of high density. We see either the
liquid or the dust phase, depending on whether we probe the system by a
moment of order $q<q_\ifmath{{\mathrm{c}}}$ or $q>q_\ifmath{{\mathrm{c}}}$, respectively.
In Fig.~4.17a, $\lambda_q$ is compiled \cite{BiaZa} from KLM, EMC and NA22
as a function of the order $q$. The low $p_\ifmath{{\mathrm{T}}}$ NA22 data \cite{Ajin89-90}
$(p_\ifmath{{\mathrm{T}}}<0.15$ GeV/$c$) indeed show a marked minimum with $q_\ifmath{{\mathrm{c}}}$ between 3
and 4, while the uncut data have not saturated at $q\leq 5$. Following
\cite{BiaZa}, the $\lambda_q$ behaviour has been studied by a number of
heavy-ion experiments \cite{Shiv93,Shiv94,Seng90,Sark93}. While a
saturation, but no clear minimum is seen by experiments stopping
their analysis at $q=5$ or 6, a minimum is now observed for $4<q_\ifmath{{\mathrm{c}}}<5$
in central C-Cu collisions at 4.5 $A$ GeV/$c$, where the analysis is carried
to $q=8$ \cite{Sark93} (Fig.~4.17b).
The observation of a minimum in the $\lambda_q$-distribution suggests a phase
transition \cite{PescTH5891,Pesc89,BrPe90}, but according to the
interpretation
\cite{BiaZa} it is merely the ``apparatus'' changing from a sensitivity
for the dominating small fluctuations at $q<q_\ifmath{{\mathrm{c}}}$ to an insensitivity for
those at $q>q_\ifmath{{\mathrm{c}}}$. The two phases could coexist without a transition being
necessary.
So, phase transition or not, two phases seem to coexist and it will be a
challenge to find
their physical interpretation in terms of the theory of strong interactions.
\subsection{Self-affinity}
Comparing log-log plots for one phase-space dimension, one notices that the
ln$F_q$ saturate, but at different $F_q$ values for different variables
$y,\varphi$ or ln$p_\ifmath{{\mathrm{T}}}$. The saturation in one dimension can be
explained as projection effect of a three-dimensional phenomenon. However,
also in three dimensional analysis the power law (2.108) is not exact. In
Fig.~4.11c, the 3D data are seen to bend upward.
It has been shown in \cite{Wu93} (see also \cite{Wosi95}) that this can
be understood by taking the anisotropy of occupied phase-space (longitudinal
phase space \cite{Hove69}) into account. In view of this phase-space
anisotropy, also its partition should be anisotropic. In other words, the
density fluctuation in phase space should be {\it self-affine} rather than
{\it self-similar} \cite{Mand91}.
If the phase-space structure is indeed self-affine, it can be characterized
by a parameter called roughness or Hurst exponent \cite{Mand91},
defined as
\begin{equation}
H_{ij}=\ln \lambda_i/\ln \lambda_j \ \ \ (0\leq H_{ij}\leq 1)
\end{equation}
with $\lambda_i$ ($i=1,2,3;\ \ \lambda_1\leq\lambda_2\leq\lambda_3$) being the shrinkage
ratios in the self-affine transformations
\begin{equation}
\delta x_i \to \delta x_i/\lambda_i\ ,
\end{equation}
of the phase-space variables $x_i$.
The Hurst exponents can be obtained from the experimentally observed
saturation curves of the one-dimensional ln$F_q(\delta x_i)$ distributions.
Using the NA22 curves for $y$ and ln$p_\ifmath{{\mathrm{T}}}$ (Fig.~4.11c), a Hurst exponent
of $H_{y,p_\ifmath{{\mathrm{T}}}}=0.516\pm 0.015$ is obtained for these two variables, in
agreement with self-affinity ($H<1$) rather than self-similarity $(H=1)$.
The upward bending for $F_q$ in the three-dimensional self-similar analysis
is then easy to understand:
Performing a self-similar analysis, phase space is not shrunk according
to the self-affine dynamical fluctuation. So, the real dynamic fluctuation
cannot be fully observed and the corresponding $F_q$ comes out smaller
at intermediate scales. At very small bins, however, this difference between
self-affine and self-similar space shrinkage disappears and the $F_q$ values
obtained approach each other. As a consequence, the slope on the log-log plot
has to increase at small bin sizes
and the self-similar analysis grants an upward bending if the
underlying structure is self-affine (i.e. corresponds to a power law).
On a self-affine Monte-Carlo branching model exactly reproducing the NA22
$d_q/d_2$ values of Fig.~4.16a, this upward-bending effect is shown to cause
the apparent violation of the L\'evy stability $\mu\leq 2$ described in
Subsect.~4.3.3 \cite{Zhan95}.
\section{Dependences of the effect}
\subsection{Charge dependence}
A mechanism known to cause correlations at small distances in phase space is
Bose-Einstein interference between identical
particles~\cite{CapFiaKrz89,CaFrie89,Gyul90}.
For reviews of the present status
of this field we refer to~\cite{Zaic,LPHEP}.
{}From the outset it must be realized, however, that the
conventional Gaussian- or exponential-type
parametrizations of the Bose-Einstein effect lead
to a saturation at $\delta y\to 0$ and {\it not} to the power law (\ref{fq113})!
In~\cite{Gyul90} it is argued that the slopes should be roughly a factor 2
larger for identical particles than for all charges combined. The experimental
situation is less than clear, in particular for 1D analyses. Contrary to
the prediction, TASSO and DELPHI see less intermittency for identical
particles. EMC finds an enhanced effect for positive but not much for
negative particles in a one-dimensional analysis, and very similar slopes
in a 3D analysis. NA22 observes an enhancement for negatives, but not for
positives. UA1 sees no difference, whereas NA35 sees an increase.
CELLO finds Bose-Einstein interference necessary to explain the residual
difference between data and JETSET 7.2, but needs an un-physically large
strength-parameter $\lambda$ to obtain agreement. In the DELPHI analysis,
Bose-Einstein interference is insufficient to explain the difference between
data and models, even with an un-physically large value of
the coherence parameter $\lambda$.
Following a suggestion in~\cite{Biya90}, higher-order
Bose-Einstein correlations have been studied
by UA1~\cite{Neum91}, NA22~\cite{Agab} and DELPHI ~\cite{Abreu95}.
In this study, ``correlation functions'' of order $q$,
\begin{equation}
R_q(Q^2_{q\pi})=N_q(Q^2_{q\pi})/N^{BG}_q(Q^2_{q\pi}),
\label{boseq}
\end{equation}
are defined as ratios of the distribution of like-charged $q$-tuplets
$(q=2,3,\dots,5)$ $N_q(Q^2_{q\pi})$ and a distribution of reference
(background) $q$-tuplets $N^{BG}_q(Q^2_{q\pi})$ obtained from random
event-mixing. The variable $Q^2_{q\pi}$ is defined as a sum over all
permutations
\begin{equation}
Q^2_{q\pi} = Q^2_{12} + Q^2_{13} + \dots Q^2_{(q-1)q}
\end{equation}
of the squared four-momentum difference $Q^2_{ij}=-(p_i-p_j)^2$ of particles
$i$ and $j$. Note that the functions (\ref{boseq}) are normalized
inclusive densities and not correlation functions in the proper sense
(cfr.~Sect.4.8).
The UA1 data are shown in Fig.~4.18. A good fit is obtained if in the
expansion
of $R_q(Q^2_{q\pi})$ suggested in~\cite{Biya90}, Gaussians (dashed curve)
are replaced by exponentials in $Q_{q\pi}$ (solid curve). Since low
$Q^2_{ij}$ pairs are lost due to limited two-track resolution in the
detector, the data at the smallest $Q^2_{ij}$ have to be regarded as a
lower limit. A power law as expected from intermittency cannot be excluded.
UA1 has further studied the distributions $R_2$ for all-charged-(cc),
$(\pm\pm)$- and $(+-)$-pairs as a function of $Q^2(\equiv Q^2_{2\pi})$
(Fig.~4.19)~\cite{Alba90}. These results have important implications. The
charge-dependence of ``intermittency", controversial in single-particle
variable analyses (see before), is now quite clear in invariant-mass variables
$(Q^2=M^2_{{\mathrm{inv}}}-4m^2_\pi)$. The data for $R^{\pm\pm}_2$ (dashed) has a much
stronger $Q^2$-dependence than $R^{+-}_2$ and effectively determines the
small-$Q^2$ behaviour of $R^{\ifmath{{\mathrm{c}}}\rc}_2$. This is unambiguous evidence that
intermittency at small $Q^2$ is predominantly due to like-sign particle
correlations. It does not necessarily imply, however, that Bose-Einstein
interference is the sole cause.
In~\cite{Arne86} it is shown on EMC data that, especially in 3D, $F^{--}_2$
deviates much more from LUND Model predictions than $F^{+-}_2$.
The LUND Model version used does not include Bose-Einstein correlations.
The deviation from the data is indicative for the importance of this effect.
Bose-Einstein interference must thus play a significant role at least for
small $Q^2$. This seems in contradiction with claimed successes in
$\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ annihilation of parton shower Monte Carlos which neglect
Bose-Einstein interference.
Finally, we reiterate our remark that a ``conventional'' Bose-Einstein effect
with exponential or Gaussian $Q$-dependence is incompatible with intermittent
power-law behaviour. We return to this point in Subsect.~4.8.4.
\subsection{Transverse-momentum dependence}
An interesting question is whether semi-hard effects~\cite{OchWo88-89},
observed to play a role in the transverse-momentum behaviour even at NA22
energies~\cite{Ajin87}, or low-$p_\ifmath{{\mathrm{T}}}$ effects ~\cite{VHov89,BialPL89} are
at the origin of intermittency. A first indication for the latter comes
from the most prominent NA22 ``spike" event~\cite{AdamPL185-87},
where 5 out of 10 tracks in the spike have $p_\ifmath{{\mathrm{T}}}<0.15$ GeV/$c$.
In Fig.~4.20a, NA22 data on $\ln F_q$ versus $-\ln\delta y$ are given for
particles
with transverse momentum $p_\ifmath{{\mathrm{T}}}$ below and above 0.15 GeV/$c$, and with
$p_\ifmath{{\mathrm{T}}}$ below and above 0.3 GeV/$c$. For particles with $p_\ifmath{{\mathrm{T}}}$ below the
cut (left), the $F_q$ exhibit a stronger $\delta y$-dependence than for
particles with $p_\ifmath{{\mathrm{T}}}$ above the cut (right).
NA22 does not claim straight lines in Fig.~4.20a, but uses fits as an
indicative measure of the increase of $\ln F_q$ over the region
$1>\delta y>0.1$. In the upper half of Fig.~4.20b, the fitted anomalous dimensions
$d_q$ are compared to those obtained in the full $p_\ifmath{{\mathrm{T}}}$-range. The
restriction to particles with $p_\ifmath{{\mathrm{T}}} < 0.15$ or $0.30$ GeV/$c$ indeed leads
to an $increase$ of $d_q$; a $decrease$ is observed for $p_\ifmath{{\mathrm{T}}} > 0.15$
or $0.30$ GeV/$c$. This observation is confirmed by IHSC~\cite{Are91}
FRITIOF predictions are given in the lower part of Fig.~4.20b,
again for all tracks and for tracks with restricted $p_\ifmath{{\mathrm{T}}}$.
It is known~\cite{Ajin89-90} that FRITIOF gives too small slopes for
factorial moments integrated over $p_\ifmath{{\mathrm{T}}}$. Here, one notices that it
also fails to reproduce their $p_\ifmath{{\mathrm{T}}}$-dependence.
UA1 has a bias against tracks with $p_\ifmath{{\mathrm{T}}}<0.15$ GeV/$c$, but
gives the dependence of $\phi_2$ on the average transverse momentum
$\bar p_\ifmath{{\mathrm{T}}}$ of the event (Fig.~4.20c) \cite{Wu94}. The data show a remarkable
decrease of $\phi_2$ with increasing $\bar p_\ifmath{{\mathrm{T}}}$ and, after passing through
a minimum at $\bar p_\ifmath{{\mathrm{T}}}\approx 0.5$ GeV/$c$, a slight increase at higher
$\bar p_\ifmath{{\mathrm{T}}}$ values. Lower $\bar p_\ifmath{{\mathrm{T}}}$ events correspond to soft processes,
while higher $\bar p_\ifmath{{\mathrm{T}}}$ ones correspond to events with hard jet
subprocesses.
Both types of events have higher slopes $\phi_2$ than their mixture at
intermediate $\bar p_\ifmath{{\mathrm{T}}}$ values. (See further in~\cite{WuLiu90} for a
possible connection to the multiplicity dependence to be described
in Subsect.~4.4.4 below.)
Fig.~4.20c also contains the results obtained from Monte-Carlo events
generated with PYTHIA~5.6~\cite{PYTHIA}. At low $\bar p_\ifmath{{\mathrm{T}}}$ values,
the PYTHIA $\phi_2$ values are strongly suppressed as compared to those
of the data.
We conclude that the intermittency observed in NA22 and UA1 data is
enhanced at low transverse momentum and is not dominated by semi-hard
effects. Hard effects dominate in high energy $\re^+\re^-$ and lh collisions.
Data on the $p_\ifmath{{\mathrm{T}}}$-dependence of factorial moments in these processes
should help in clarifying the origin of intermittency. The effect of
$p_\ifmath{{\mathrm{T}}}$-cuts on $\re^+\re^-$ data has been studied by DELPHI~\cite{Abreu90}.
One-dimensional data are shown in Fig.~4.21 and provide
several important pieces of information:
i) The log-log plot for low-$p_\ifmath{{\mathrm{T}}}$ particles shows less saturation (i.e.
stronger intermittency) than for larger $p_\ifmath{{\mathrm{T}}}$ particles. So, intermittency
is strongest in the $p_\ifmath{{\mathrm{T}}}$-region where hard gluon effects are weakest!
ii) A discrepancy between data and models (only indicative in Fig.~4.7d above)
is observed in the interval $0.255<p_\ifmath{{\mathrm{T}}}<0.532$ GeV/$c$. This looks
surprising at first, but we shall show in Subsect.~4.4.4 that the
intermittency
effect can be stronger for individual mechanisms than for a mixture.
iii) The factorial moments are larger for $p_\ifmath{{\mathrm{T}}}>0.532$ GeV/$c$ than for
$p_\ifmath{{\mathrm{T}}}<0.255$ GeV/$c$, opposite to the trend of the NA22 data (Fig.~4.20a).
Also this seems contradictory, but it should be realized that for NA22
transverse momentum refers to the beam axis, which is usually close to beam
and target jet-axes. In the $\re^+\re^-$ analysis, $p_\ifmath{{\mathrm{T}}}$ is calculated relative to
the global event axis which differs from the direction of individual jets.
\subsection{Dependence on jet topology}
In their recent analysis, DELPHI~\cite{Abreu90} selects 2-jet and 3-jet events
using the {JADE/E0} invariant-mass algorithm~\cite{Jade:clus}, with resolution
parameter values $y_{{\mathrm{cut}}}=0.04$ and 0.01, and with additional cuts to clean
the 2-jet and 3-jet sample. At large bin sizes, factorial moments rise
faster with decreasing bin size (and are, therefore, larger) in 3-jet
than in 2-jet events.
This is compatible with the (large bin-size) behaviour expected from hard
gluons. At small bin sizes the increase is similar for 2-jet and 3-jet events.
In 3-jet events, factorial moments were calculated for tracks belonging
to jet 1, jet 2 and jet 3 ordered in energy. The rapidity was defined with
respect to the individual jet axis. As seen in Fig.~4.22, intermittency is
weakest in jet~3 and strongest in jet~2. The deviation from JETSET is
also strongest for jet~2.
\subsection{Multiplicity (density) dependence}
In general, a decrease of the intermittency indices $\phi_q$ is found with
increasing energy, in particular for hh, h$A$ and $AA$ collisions.
As seen in Fig.~4.23a, a strong multiplicity dependence of the intermittency
strength is observed for hh collisions by UA1~\cite{Alba90}. The trend is
opposite to the predictions of the models used by this collaboration. This
decrease of the intermittency strength with increasing multiplicity is usually
explained as a consequence of mixing of independent sources of
particles~\cite{LiBu89}. The cross-over of data and FRITIOF at intermediate
multiplicity explains the apparent success of FRITIOF in Fig.~3.3,
for multiplicities close to 30 as being accidental.
Mixing of emission sources leads to a roughly linear decrease of the
slopes $\phi_q$ with increasing particle density $\langle \rho\rangle$ in
rapidity~\cite{CapFiaKrz89,BiaFest89,Seib90}: $\phi_q\propto \langle\rho\rangle^{-1}$.
This is indeed observed by UA1~\cite{Alba90}.
Multiple emission sources are present in multi-chain Dual Parton models.
The calculated slopes indeed depend linearly on multiplicity but are
too small by a factor of two~\cite{Aur91}. Similarly, the model studied
in~\cite{Barsh90} with independent emission at fixed impact-parameter
finds decreasing $\phi_q$ with increasing multiplicity.
Also here, a study of the multiplicity dependence in $\re^+\re^-$ data and JETSET
allows interesting comparisons. In fact, the LEP results~\cite{Abreu90}
suggest little or no $n$-dependence, except for the lowest multiplicities,
where the slope is largest and also the difference with JETSET~PS is the
largest.
Fig.~4.23a helps in explaining why intermittency is so weak in heavy-ion
collisions (cfr.~Fig.~4.3): the density (and mixing of sources) is
particularly high there.
In Fig.~4.23b EMU01~\cite{Adamo90}, therefore, compares $\phi_2$ for NA22 (hp
at 250 GeV) and heavy-ion collisions at similar beam momentum per nucleon,
as a function of the particle density. Whereas slopes averaged over
multiplicity are smaller for $AA$ collisions than for NA22 in Fig.~4.3, at
fixed $\langle \rho\rangle$ they are actually higher than expected from an
extrapolation of hh collisions to high density and even grow with
increasing size of the nuclei. This may be evidence for re-scattering
(see~\cite{Verlu90}) or another (collective) effect, but, as shown by
HELIOS~\cite{AAke90} and recently confirmed by EMU-01~\cite{Adamo90}, one has
to be very sure about the exclusion of $\gamma$-conversions before drawing
definite conclusions.
We conclude this section with an additional warning. In Subsect.~4.3.2 we
mentioned the Fia\l kowski ``universality-conjecture'' and noted that it is
incompatible with the ``mixing'' hypothesis usually invoked to explain the
multiplicity dependence of factorial moments and slopes. A different
explanation of the multiplicity dependence may therefore be needed, especially
since intermittency and Bose-Einstein effects are now known to be closely
related.
\section{Factorial cumulants}
Normalized factorial cumulant moments, first introduced in~\cite{Mue71} and
recently studied in~\cite{CaES91}, are defined in (\ref{dr:47}) as integrals
over the background subtracted correlation functions. They share with
factorial moments the property of ``noise suppression''. The normalized
factorial moments $F_q$ can be expanded in terms of normalized cumulant
moments $K_q$ as given in (\ref{dr:48}). This expansion has been found to
converge rapidly~\cite{CaES91}. The terms in the expansion correspond to
contributions from genuine $q,(q-1)\dots,2$-particle correlations.
An analysis of factorial cumulant moments is presented in~\cite{CaES91}.
Roughly, it is estimated that $K_2\sim0.6, K_3\sim 0.7, K_4\simkl1.0,
K_5\simkl1.5$ for UA1 data at $\sqrt s=630$ GeV. (The inequalities for $K_4$
and $K_5$ are due to the approximation $\overline{AB}$ by $\bar A\cdot\bar B$ in
(\ref{dr:48}) since no direct measurements of these averages exist.)
Clearly, the two-particle contribution to factorial moments is large,
but higher orders are not negligible. At the energy of the NA22 experiment
$K_2$ is small ($\sim$0.2), but $K_3$ is significantly larger ($\sim$0.45).
{}From (\ref{dr:48}) it is seen that the contribution $F^{(2)}_q$ to $F_q$
from
two-particle correlations alone can be expressed as
\begin{equation}
F^{(2)}_3=1+3K_2\nonumber
\end{equation}
\begin{equation}
F^{(2)}_4=1+6K_2+3\overline{K^2_2}\ \ ;
\end{equation}
the contribution $F^{(3)}_q$ from two- and three-particle correlations
to $F_4$ as
\begin{equation}
F^{(3)}_4=1+6K_2+3\overline{K^2_2}+4K_3\ \ .
\end{equation}
The difference $F_q-F_q^{(p)}$ is a measure for the importance
of higher-order correlations.
Fig.~4.24a shows a cumulant-decomposition of $F_3$ and $F_4$ in
UA1-data~\cite{Egg91}. The differences between the curves indeed indicate
large contributions from genuine higher-order correlations. Similar results
are observed for NA22~\cite{Ajin89-90} in Fig.~4.24b, for $p=$2 and 3 and
$q=$3 and 4, in one-, two- and three-dimensional phase space (transformed
$y, y-\varphi$ and $y-\varphi-\ln p_\ifmath{{\mathrm{T}}})$. In general, the difference increases with
increasing ln$M$ (decreasing bin size). This means that the contribution of
higher-order correlations to the factorial moments increases at higher
resolution. An exception are factorial moments in the variable $\varphi$,
for which only two-particle correlations are found to be non-zero (not
shown).\footnote{Absence of genuine higher correlations has been reported
in~\cite{Wang94-2}, but at far too low statistics.}
The situation is completely different in heavy-ion collisions where, with
present accuracy, $K_q\approx 0$ for $q>2$ (Fig.~4.24c). The factorial
moments are completely dominated by two-particle
correlations~\cite{Egg91,JaMuSi,Adam93},
implying that higher-order $F_q$ contain little or no further dynamical
information for this type of collisions (see~Eq.~\ref{dr:37a}).
Using the linked-pair ansatz~\cite{CaES91} (see further Sect.~5.1.1.),
higher-order cumulant functions can be expressed as products of $K_2$ (see
also~\cite{Dia90} for an interpretation in terms of
independent superposition of sources)
\begin{equation}
K_q = A_q K_2^{q-1},
\end{equation}
with free constants $A_q$.
For a negative-binomial (NB) multiplicity distribution, $K_2=1/k$ and the
linking parameters are fixed numbers given by
$A^{NB}_q=(q-1)!$~\cite{WolfAPP90}. A necessary condition is stationarity,
i.e. constancy of $1/k$. This works well for UA1. For NA22~\cite{Ajin89-90},
$A_q$ is observed to increase with decreasing bin size. Approximately constant
$A_q\approx (q-1)!$ are found if the data are averaged only over a narrow
rapidity region $(-0.75\leq y\leq0.75)$ and the most prominent spike event is
excluded. The linked-pair ansatz may thus be a valid approximation for
high-order correlations in small phase-space domains but not for the average
over phase space. This would be consistent with the well-documented
fact~\cite{GiovHove86} that the negative binomial is often a good
parametrization of multiplicity distributions in restricted $\delta
y$-intervals.
We shall come back to cumulants and genuine higher-order correlations in
Sect.~4.10, where they are studied by means of a largely improved
methodology.
\section{Factorial correlators}
\subsection{The method}
The moments defined
in (\ref{dr:44}-\ref{dr:46}) measure local density
fluctuations in
phase space. Additional information is contained in the correlation between
these fluctuations within an event. This correlation can be studied
by means of the factorial correlators defined in (\ref{f4:2}).
Correlators are
typically calculated at a given $\delta y$ for each combination $mm'$ of bins
with size $\delta y$, and then
averaged over all combinations separated by a given bin-distance $D$.
This is illustrated below.
\vspace*{1.3truecm}
\epsffile[20 60 80 100]{raster.ps}
\vskip 2mm
In the simple intermittency model ($\alpha$-model) described
in~\cite{bialas1,bialas2}, $F_{pq}$ depends on
$D$ but not on $\delta y$ and follows the power law
\begin{equation}
F_{pq} \propto (\Delta Y/D)^{\phi_{pq}}\ \ .
\label{3.27n}
\end{equation}
\noindent
The powers $\phi_{pq}$ (slopes in a log-log plot) obey
the relations~\cite{bialas1,bialas2}:
\begin{equation}
\phi_{pq} = \phi_{p+q} - \phi_p - \phi_q = pq\phi_2\ \ ,
\label{3.28n}
\end{equation}
\noindent
where the first equality sign is due to the $\alpha$-model proper,
the second to the
log-normal approximation. According to (\ref{3.27n}) $\phi_{11}=\phi_2$,
so that (\ref{3.28n}) can also be written in the form
\begin{equation}
\phi_{pq} = pq \phi_{11}\ .\end{equation}
\subsection{Results}
Preliminary results for pseudorapidity resolution $\delta \eta\geq 0.2$ have been
reported by the HELIOS Collaboration~\cite{HELIOS89}. There, however,
multiplicities $n_m$ had to be estimated from the transverse energy
$E_{\ifmath{{\mathrm{T}}},m}$ in bin $m$ and the average transverse energy $\langle E_\ifmath{{\mathrm{T}}}\rangle$
per particle: $n_m=E_{\ifmath{{\mathrm{T}}},m}/\langle E_\ifmath{{\mathrm{T}}}\rangle$ rounded to the nearest integer.
The first direct measurement is from NA22~\cite{Aiv91}. The $\ln F_{pq}$
are shown as a function of $-\ln D$ in Fig.~4.25a-d, for four values of
$\delta y\geq 0.1$ (corresponding to $M=10$, 20, 30 and 40). Statistical errors
(estimated from the dispersion of the $F_{pq}$ distribution) are in general
smaller than the size of the symbols. $F_{pq}$ can be measured
up to third order in $p$ and $q$ for $\delta y$=0.4 binning (Fig.~4.25a).
For $\delta y$=0.1, the analysis is possible to first and second order only
(Fig.~4.25d). The smallest possible value for $D$ being equal to the bin size
$\delta y$, Fig.~4.25a extends to $D=0.4$ and Fig.~4.25d to $D=0.1$. In all
cases,
an increase of $\ln F_{pq}$ is observed with increasing $-\ln D$. Very
similar results have recently been reported by EMC~\cite{Dera90},
EMU-01~\cite{Adamo90} and in \cite{Ghosh92}.
In Fig.~4.26a, the $\ln F_{pq}$ are compared at fixed $D=0.4$ for four
different
values of $\delta y$. The dashed lines correspond to a horizontal line fit through
the data. In agreement with the $\alpha$-model, the $F_{pq}$ indeed do not depend
on $\delta y$. Also this result has been confirmed on EMC data~\cite{Dera90}
and in \cite{Ghosh92}.
The $\delta y$-independence of correlators holds exactly in the $\alpha$-model.
Nevertheless, Fig.~4.26b shows that the $\delta y$ independence is also valid in
FRITIOF. For the particular value of $D=0.4$, this even happens at very
similar values of $\ln F_{pq}$ as in the data. In fact, this property is
far from unique to the $\alpha$-model, but holds approximately in any model with
short-range order~\cite{dewolf92}.
For $F_{11}$, the $\delta y$ independence is easily derived from a parametrization
of the two-particle density, integrated over two regions of size $\delta y$
separated by $D$. Using exponential short-range order~\cite{WolfAPP90}, this
gives
\begin{equation}
F_{11} -1~~ \propto~~\frac{1}{a^2}
\ifmath{{\mathrm{e}}}^{-D/L}/ (\ifmath{{\mathrm{e}}}^a-1)(1-\ifmath{{\mathrm{e}}}^{-a}) \ \ ,
\label{3.29}
\end{equation}
\noindent
where $L$ is a correlation length and $a=\delta y/L$. According to (\ref{3.29}),
$F_{11}$ becomes independent of $\delta y$ for $a\ll1$. Since $\ifmath{{\mathrm{e}}}^{-D/L}\to 1$
as $D\to 0$, this form also leads to the deviations from (\ref{3.27n})
observed as a bending in Fig.~4.25.
Because of the bending, fitted slopes $\phi_{pq}$ have no meaning, except
as an indication for the increase of $F_{pq}$ in a restricted range.
The slopes for two values of $\delta y$ are compared to FRITIOF predictions
in Fig.~4.27a and 4.27b, respectively. As observed earlier for the case
of univariate moments~\cite{Ajin89-90}, the FRITIOF slopes are too small
also for the correlators. This is not surprising since the model does not
succeed in reproducing even the lowest-order (i.e.~two-particle)
rapidity correlation function (Chapter~3).
\subsection{Interpretation}
Factorial correlators have been analysed in~\cite{eggers:correlators}
using a suitable parametrization of\break
$K_2(y_1,y_2)$ and the linked-pair
ansatz~\cite{CaSa89} for higher-order correlations. The relations
(\ref{eq:b7}-\ref{eq:b10}) of Sect.~2.1.6 then allow to express all
correlators in terms of $K_2$ for arbitrary ($p,q$). Note that the
expressions for $F_{pq}$ contain many lower-order ``combinatorial''
terms which effectively dominate and mask the contribution from
genuine ($p+q$)-order correlations.
A basically similar analysis is presented in~\cite{dewolf92}, inspired
by techniques used in quantum-optics. The two analyses have no
difficulty to describe basic features of the NA22 data, including the
sum-rules discussed in Sect.~2.3 which were claimed to be a unique
test of random cascade models. Fig.~4.28a compares NA22 data on
$F_2(\delta y)$ and $F_{11}(D)$ with the calculations
from~\cite{dewolf92}. $F_2(\delta y)$ is used to fix the parameters of
$K_2(\delta y)$ (assuming stationarity); $F_{11}(D)$ follows after
integration over the appropriate rapidity-domains. With the
linking-ansatz of~\cite{WolfAPP90} all other correlators are
calculated without further assumptions. An illustrative example is
shown in Fig.~4.28b which compares $F_{12}(D)$ from NA22 to the
prediction. The agreement is excellent in all cases. This observation
is confirmed in \cite{Ghosh92}.
According to (\ref{3.28n}), the ratio $\phi_{pq}/\phi_2$ is expected to grow with
increasing orders $p$ and $q$ like their product $pq$. In Figs.~4.27c and
4.27d
this is tested for $\delta y$=0.4 and $\delta y$=0.2, respectively. In both cases,
the experimental results lie far above the dashed line corresponding to the
expected $\phi_{pq}/\phi_2=pq$. Since the dependence of $\ln F_{pq}$ on $-\ln D$
is not strictly linear, this comparison depends on the range of $\delta y$ and
$D$ used to determine $\phi_2$ and $\phi_{pq}$. In Fig.~4.27d one, therefore,
compares a number of fits. Slopes are smaller when the upper limit in $D$
is reduced, but do not reach the $\alpha$-model prediction (dashed line).
It can be verified that, at least for the higher orders, the
discrepancy with (\ref{3.28n}) is mainly due to the second equal sign,
derived from a log-normal approximation to the density
distribution. In a recent paper~\cite{AlbBi91}, this approximation has
been shown to be valid if the density fluctuations are weak or if the
density probability distribution is log-normal. The NA22 data
demonstrate that none of these conditions is fulfilled.
We conclude that the correlators $F_{pq}$ increase with decreasing
correlation length $D$, but do not really follow a power law for
$D\ ^<\hs-2.5mm_\sim\ 1$. For fixed $D$, the values of $F_{pq}$ do not depend on the
resolution $\delta y$, a feature expected from the $\alpha$-model, but also
reproduced by FRITIOF and approximately true in any model with
short-range order. When the increase of the correlators is roughly
approximated by a straight line in a restricted interval, the powers
$\phi_{pq}$ increase linearly with the product $pq$ of the orders, but
are considerably larger than expected from FRITIOF and from the simple
$\alpha$-model.
The extension of single-variate factorial moments to the multivariate
case offers better insight into the complicated nature of the
correlations. However, the original expectation that correlators would
help in clarifying the issue of intermittency is not borne out by
present data. Simple but reasonable models for higher-order
correlation functions which use the experimental 2-particle
correlations as input, have no difficulty in reproducing the behaviour
of factorial correlators measured e.g.~by NA22.
\section{Multifractal analysis}
Power-law dependence of normalized factorial moments on the
resolution $\delta$ (bin size) is a signature of self-similarity in the
fluctuation pattern of particle multiplicity. It suggests
that the probability distribution $P(\rho,\delta)$ of the
particle density $\rho$ has fractal properties.
For simple Widom-Wilson~\cite{Widom65} type scaling,
$P(\rho,\delta)$ is of the form
\begin{equation}
P(\rho,\delta)\sim\delta^{-\beta} P^{\star}(\rho/\delta^\nu)\ \ ,
\label{3:eq:1}
\end{equation}
where $\beta$ and $\nu$ are critical exponents.
All $q$-th order moments of $\rho$ ($q=1,2,\ldots$) obey
power laws in $\delta$ with inter-related exponents depending
on $q$ and on ($\beta$, $\nu$). This characterizes a simple or
mono-fractal. Another possibility is a multifractal behaviour, in which
$P(\rho,\delta)$ obeys a relation of the type~(cfr.~\cite{Kadanoff89})
\begin{equation}
\ln P(\rho,\delta)/\ln\delta =f(\alpha),
\quad \alpha=\ln\rho/\ln\delta.\label{3:eq:2}
\end{equation}
Multifractals, first introduced in~\cite{Mandelbrot74} represent
infinite sets of exponents---the multifractal spectrum---which
describe the power-law scaling of all moments of
$P(\rho,\delta)$. In principle, knowledge of the multifractal spectrum
is completely equivalent to knowledge of the probability distribution.
Unlike geometrical or statistical systems, multiparticle production
processes pose special problems if a multifractal analysis is
to be considered. The most obvious one is the finiteness of particle
multiplicity in an event at finite energy.
Self-similarity, if existent, therefore cannot persist indefinitely to
finer and finer scales of resolution.
In multiparticle production $P(\rho,\delta)$ is not directly accessible.
At best one can construct, for a single event of multiplicity
$n$ and for given $\delta$, a frequency distribution
which approaches $P(\rho,\delta)$ only for $n\rightarrow\infty$.
For any finite (and usually small) $n$, the frequency distribution
and its moments will be subject to statistical fluctuations.
Since the data sample contains a large number of events, it is obviously
recommended to consider the event average. This averaging, however,
supposes ergodicity. The applicability of the multifractality concept
can, therefore, only be justified a posteriori.
\subsection{The method }
A multifractal analysis is based on the the properties
of $G$-moments whose definition is given in ~(\ref{gghwa}).
The moments $G_q$ (or more often $\ln G_q$) are obtained
for each individual event at a specified resolution $\delta y\sim1/M$ and
then averaged over the event sample\footnote{
Note that analyses based on $\langle G_q\rangle$ or on $\langle \ln G_q\rangle$
in general differ and probe different aspects of the system
under study~\cite{Aharony89}.}.
In the theory of multifractals, the $G$-moments share
with the scaled factorial moments the property that self-similar density
fluctuations lead, in principle, to scaling behaviour
\begin{equation}
G_q \propto (\delta y)^{\tau_q} \ \ \ \ {\mbox{for}}\ \delta y\to 0\ .
\end{equation}
In a fractal analysis (see also Sect.~2.4), one therefore determines
the slope
\begin{equation}
\tau(q,M) = - \frac{\partial\langle \ln G_q(M)\rangle}{\partial\ln M} \end{equation}
on a double-logarithmic plot, after averaging over all events in the
sample.
A multifractal spectral function is introduced via a Legendre
transform defined as
\begin{equation}
f(\alpha_q) = q\,\alpha_q - \tau_q \ \ , \end{equation}
with
\begin{equation}
\alpha_q=\frac{\partial\tau_q}{\partial q} \end{equation}
being the Lipschitz-H\"older exponents.
The spectral function $f(\alpha_q)$ is a smooth function, concave downward,
with its maximum at $q=0$. It gives a quantitative description of the
density fluctuations in the dense and in the sparse regions, corresponding to
its left and right wing, respectively. A wide spectrum reveals a non-smooth
density distribution.
The generalized (R\'enyi)-dimensions are given by
\begin{equation}
D_q = \frac{1}{q-1}\, \tau_q=\frac{1}{q-1}\, \left[q \alpha_q-f(\alpha_q)\right]\ . \end{equation}
\subsection{Experimental results}
$G$-moments have been studied in a number of experiments~
\cite{Dera90,Ghosh92,Shiv93,Shiv94,Seng90,Ghosh91,Sark93,SugG,EMCG,IHCG}
\cite{NA22G,UA1G,CDFG,Wang95,Sark93-2,Shiv93-2,Jain90}.
As an illustrative example, we show UA1 results~\cite{UA1G} in Fig.~4.29a,
where the event average $\langle \ln G_q\rangle$ is plotted as a function of the
resolution $(M=2^\mu)$. Starting at a value of $0$ for $\mu=0$
according to definition (\ref{gghwa}), the moments grow
for $q<1$ and fall for $q>1$ as $\mu$ increases. The slopes decrease and
$\langle\ln G_q\rangle$ tends to saturate for large $\mu$. The saturation is due
to
an increasing number of bins with content $n_m=0$ or 1, as $M$ becomes large,
\begin{equation}
G_q(M) \to n\left( \frac{1}{n}\right)^q = n^{1-q} {\rm \ \ \ \ for\ }
M\to\infty\ \ . \end{equation}
Fig.~4.29b shows $\langle\tau_q\rangle$ and $\langle\alpha_q\rangle$ for $\mu=1$ and 2
(small $M$). The corresponding spectral function $\langle f(\alpha_q)\rangle$ is given
in Fig.~4.29c as a function of $\langle \alpha_q\rangle$.
The fact that $\langle f(\alpha_q)\rangle$ does not degenerate into a single point
implies multifractality in hadron production, at least for large bin
size $\delta y=\frac{\Delta Y}{M}$ (small $\mu$). However, for smaller bin sizes,
the function turns over (i.e., bends upward) and falls into the
non-physical region above the dashed line (not shown).
\subsection{Universality}
{}From Fig.~4.29c, it is clear that $\langle f(\alpha)\rangle$
depends on $\mu$.
It also depends on the cms energy $\sqrt s$ (or the multiplicity
$n=2^\nu$).
In~\cite{Flor91} it is
conjectured that $G$-moments (at fixed $q$) show universality in
$\xi=\mu-\nu$, however. The latter quantity is directly related
to the average particle multiplicity per bin,
$n/M=2^{\nu-\mu}=2^{-\xi}$.
Using a branching model, the authors of~\cite{Flor91} derive
the universality relation
\begin{equation}
\Gamma_q(\xi) = \ln G_q(\mu,\nu)-\ln G_q(\nu,\nu).\label{gmunu}
\end{equation}
It expresses the scaling behaviour of a function of two variables
($\mu,\nu$) in terms of a function of one variable only.
The function $\Gamma_q(\xi)$ determines the $G$-moments as functions of
$\mu$ for all values of $\nu$.
The validity of (\ref{gmunu}) is claimed to be a strong evidence for
self-similarity.
Fig.~4.30a shows $\Gamma_q$ as a
function of $\xi$ for $q=\pm5$. All data points are close to universal
lines, thus indeed indicating universal behaviour.
A further prediction is that also $\langle f(\alpha_q)\rangle$ is universal
for fixed $\xi$. Fig.~4.30b demonstrates that this is not confirmed
by the UA1 data~\cite{UA1G}. In the EMC data~\cite{Dera90} the left
branch shows universality, but not the right one.
Besides being more sensitive to universality breaking than $\Gamma_q(\xi)$,
the function $\langle f(\alpha_q)\rangle$ also reveals more clearly shortcomings in
the models. For PYTHIA and GENCL this is illustrated by Fig.~4.30b.
\subsection{Modified G-moments}
As stated earlier, $G$-moments have the advantage that not only spikes
are included in the analysis, but also non-empty valleys (for $q<0$).
Disadvantages are that the moments saturate
at $\delta y\to 0$ when the content of non-empty bins approaches unity and that
statistical fluctuations are not filtered out
(see also~\cite{ChiuFialHwa90,Seibert}).
In~\cite{HwaPan} a modified definition of the $G$-moments
is proposed in an attempt to circumvent the problem of statistical noise.
``Truncated'' $G$-moments are defined as
\begin{equation}
G_q = \sum^M_{m=1} p^q_m \Theta (n_m-q)\ ,
\label{ggmodif}
\end{equation}
where $\Theta$ is the usual step-function equal to 1 for $n_m\geq q$ and zero
otherwise. For very large multiplicity $n$ (as in a macroscopic statistical
system), $n/M\gg q$ and (\ref{ggmodif}) is in practice identical to
(\ref{gghwa}). In particle physics, $n$ is a relatively small number and the
$\Theta$ function exerts a crucial influence on the $G$-moments. It imposes
non-analytical cut-offs at positive integer values of $q$. With the help of a
Monte Carlo (ECCO) based on the Geometrical Branching Model, the authors show
that $\ln\langle G_q\rangle$ now exhibits a linear dependence on ln$M$ for $q>1$,
without saturation.
For $q>1$, the linearity of $\ln\langle G_q\rangle$ with $\ln M$ has
been verified on $\mu$p, $\Pap\Pp$ and $\re^+\re^-$ data~\cite{EMCG,UA1G}.
The slopes $\tau_q$ turn out to be very similar in all three reactions
and roughly equal to $\tau_q=-0.9(q-1)$.
\subsection{Bernoulli trials and $G$-moments}
Before we conclude the discussion of experimental characteristics
of $G$-moments, it is of interest to inquire in more detail about the
dynamical content revealed in multifractal analyses.
This is best done in a comparison to a model without dynamics.
Let $P(n)$ be the probability distribution for observing $n$ particles
in an initial wide interval $\Delta Y$. Let this interval be
subdivided into $M$ smaller intervals of size $\delta y=\Delta Y/M$, each
of which contains $n_m$ particles ($n_m=0,1,\ldots,n$; $m=1,2,\ldots,M$) with
$\sum_m n_m=n$.
Assume that for every subdivision of $\Delta Y$ the $n$ particles
are independently distributed over the intervals with probability $1/M$.
For fixed $n$, the joint occupation probability in $M$ cells is
given by~(see also~\cite{ChiuFialHwa90})
\begin{equation}
P_n(n_1,\ldots,n_M)=P(n)\,\frac{n!}{n_1!\ldots n_M!}\,
\left(\frac{1}{M}\right)^n \,\delta(\sum_j n_j -n).\label{3:eq:3}
\end{equation}
With (\ref{3:eq:3}) the $G$-moments (\ref{gghwa}) at fixed $n$ are given by
\begin{equation}
G_q(n,M)= M\,n^{-q}\,\sum_{i=1}^n i^q \,B(n,1/M;i),\label{3:eq:4}
\end{equation}
where $B(n,1/M;n)$ is the binomial distribution.
For integer $q\geq1$ one obtains
\begin{equation}
G_q(n,M)= M\,n^{-q}\,\sum_{j=1}^q {\cal S}_q^{(j)}\,n^{[j]}. \label{3:eq:5}
\end{equation}
${\cal S}_q^{(j)}$ is a Stirling number of the second kind
(cfr.~(\ref{dr:43})) and $n^{[j]}=n(n-1)\ldots(n-j+1)$.
In the Poisson limit of the binomial ($n$ large and $1/M$ small with
$n/M$ fixed), (\ref{3:eq:5}) simplifies further to
\begin{equation}
G_q(n,M)= M\,n^{-q}\,\sum_{j=1}^q {\cal S}_q^{(j)}\,(n/M)^j. \label{3:eq:6}
\end{equation}
In inclusive analyses the average over $P(n)$ has to be taken in
(\ref{3:eq:5}-\ref{3:eq:6}). This introduces the (inverse) moments
$\langle n^{j-q}\rangle$ and $\langle n^{[j]}/n^q\rangle$ of the multiplicity
distribution in $\Delta Y$.
Numerical studies indicate that (\ref{3:eq:4}-\ref{3:eq:6}) reproduce and
explain many of the multifractal and universality properties
seen in the data. Here we can only give a few examples.
With respect to the structure of (\ref{3:eq:4}) it should be noted that
the binomial distribution is in fact a multifractal\footnote{This is
easily verified using Stirling's approximation to $n!$ which is notoriously
accurate even for quite small $n$.} in the sense of (\ref{3:eq:2})
\cite{Billingsley65,Kadanoff89}. Consequently, ``proper'', but quite trivial
binomial multifractal behaviour will be seen if the data are noise-dominated.
This is the case in practically all analyses referred to before. This point
was recognized in~\cite{ChiuFialHwa90}, but its full consequences were not
further studied.
{}From (\ref{3:eq:6}) follows immediately that the function $\Gamma_q(\mu,\nu)$
defined in (\ref{gmunu}) depends only on the ratio $n/M=2^{-\xi}$. Thus, the
parameter-free function $\Gamma_q(\mu,\nu)$ is indeed, but trivially,
universal in the Poisson limit. It describes accurately the data in
Fig~4.30a. Being a purely mathematical property of noise, it is not
surprising that the usual Monte-Carlo models also show this type of
universality. $\Gamma_q(\mu,\nu)$ ceases to be universal in the (more
general) binomial case, although the deviations remain small in cases of
practical interest. This probably explains the universality breaking observed
in $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ Monte-Carlo simulations at $1-10$ TeV in~\cite{Chiu91}.
The above considerations can be extended to the modified $G$-moments defined
in (\ref{ggmodif}). In particular $\Gamma_q(\mu,\nu)$ remains universal in
the Poisson limit. Further numerical properties of modified $G$-moments are
illustrated on Fig.~4.31(a-c). The results shown are based on (\ref{3:eq:6})
further averaged over $n$ with a negative binomial distribution truncated at
$0$. They depend only on $\langle n\rangle$ in $\Delta Y$ and on the NBD parameter
$k$.
Fig.~4.31a demonstrates that the modified $G$-moments can be well
approximated
by power laws. The pseudo-linearity extends over a much larger interval
in $M$ than for usual $G$-moments obtained from (\ref{3:eq:6}). The improved
linearity is due to negative terms in the expressions of truncated moments
of the Poisson (or binomial) distribution. The calculations displayed in
this figure coincide (up to an overall normalization factor) nearly exactly
with the EMC data for $M\geq8$ shown in~\cite{EMCG}. This proves that the
``clear asymptotic power-law behaviour of $\langle G_q\rangle$ characteristic for
a self-similar system'' \cite{EMCG} is in fact due to Bernoulli noise.
Fig.~4.31b shows the ``Ochs-Wo\v siek'' plot for modified $G$-moments. Here
again, the quasi-perfect linear relation between $\ln G_q$ and $\ln G_{2}$
is seen to be a characteristic of Bernoulli trials. This linearity property
holds in fact for any combination $\ln G_q$ and $\ln G_{q'}$. The exponents
$\tau_q$ derived from power-law fits to the ``data points'' in Fig.4.31a are
shown in Fig.~4.31c. The line is a fit with the form $\tau_q=-C(q-1)$.
The slope $C$ is a slowly-changing function of the NBD-parameter $k$ with a
value around $0.9$, as experimentally observed in Subsect.~4.7.4 above!
\subsection{Evaluation of noise and connection between $F_q$ and $G_q$}
The self-similar property of multiparticle production at high energy can, in
principle, be investigated by $F$-moments and by $G$-moments. The power-law
behaviour of the scaled $F$-moments provides evidence for a self-similar
cascading process of dynamical origin. The $G$-moments, as an ingredient
of fractal theory, are designed to describe the multifractality aspect of
high multiplicities. In the real environment of high energy collisions,
however, the multiplicities are rather low and the $G$-moments
are dominated by statistical fluctuations.
The $F$-moments are defined for integer powers $q\geq 1$, the $G$-moments
for all real powers $q$. In order to establish a connection to
the $F$-moments, the powers $q$ are restricted to integer values of $q\geq 1$
here also for the $G$-moments.
The number $n_m$ of particles per subdivision $\delta y=\Delta y/M$ has to be equal
to, or larger than $q\ (n_m=q+k,k=0,1,\dots)$ for the $F$- and $G$-moments.
Functions
\begin{equation}
B_{q,k}(M) = \langle\frac{Q_{q+k}(M,n)}{n^q}\rangle
\end{equation}
are defined from the number of bins $Q_{n_m}(M,n)$ containing
$n_m=q+k$ particles in an event of multiplicity $n$ in the total phase-space
region $\Delta y$, normalized by $n^q$ and averaged over all events.
They express the basic fractal structure of the data, if they show
a power-law behaviour of the form
\begin{equation}
B_{q,k}(M) \propto M^{\lambda_{q,k}(M)}\ \ .
\end{equation}
In order to
suppress statistical fluctuations, the $G$-moments can be defined as the
event average over (4.33), or, equivalently, as
\begin{equation}
\langle G_q(M)\rangle = \sum^\infty_{k=0} B_{q,k}(M)(q+k)^q\ \ .
\end{equation}
They are proportional to $M^{-\tau_q}$ for large $M$.
The $F$-moments are defined as
\begin{equation}
\langle F_q(M)\rangle = M^{-1} \sum^M_{m=1} \langle \frac{(n_m(n_m-1)\dots
(n_m-q+1)}{(n/M)^q}\rangle
\end{equation}
or, equivalently, as
\begin{equation}
\langle F_q(M)\rangle = M^{q-1} \sum^\infty_{k=0} B_{q,k}(M)\frac{(q+k)!}{k!}\ \ .
\end{equation}
They are proportional to $M^{\phi_q}$ for large $M$
(note, however, that (4.41) is different from the form (2.68)
of the $F$-moments generally used).
When (4.33) and (4.40) to (4.42) are applied to the data they should show
a power-law behaviour for large $M$, if there are fractal structures present
in the data.
The dynamical contribution to the $G$-moments can be expressed by
\begin{equation}
\langle G_q\rangle^{{\mathrm{dyn}}} = \frac{\langle G_q\rangle}{\langle G_q\rangle^{{\mathrm{st}}}} M^{(1-q)}\ \ ,
\end{equation}
where $\langle G_q\rangle^{{\mathrm{st}}}$ can be determined by distributing the $n$
particles of an event randomly in $\Delta y$.
The randomization procedure destroys short-range particle correlations, but
does not alter the Bernoulli nature of the particle repartition in smaller
bins
discussed in Subsect.~4.7.5. As a result, this method does not eliminate the
binomial, noise-induced multifractal behaviour, but just gives its behaviour.
When $\langle G_q\rangle^{{\mathrm{st}}}$ is equal to $\langle G_q\rangle$, a trivial ``dynamical''
effect remains: a flat $dn/dy$ leads to a probability $1/M$ for a particle
to be in a given bin and $\langle G_q\rangle^{{\mathrm{dyn}}}=M^{1-q}$.
The dynamical contributions to the slope $\tau_q$ can be expressed by
\begin{equation}
\tau^{{\mathrm{dyn}}}_q = \tau_q-\tau^{{\mathrm{st}}}_q + q - 1\ \ ,
\end{equation}
where $\tau^{st}_q$ is the statistical part of the slope.
Subtracting the statistical contribution from $\tau_q$ gives
\begin{equation}
\tau_q-\tau^{{\mathrm{st}}}_q = \tau^{{\mathrm{dyn}}}_q - q + 1 \approx -\phi_q
\end{equation}
which can be directly compared to the slopes $\phi_q$ obtained from the
$F$-moments \cite{HwaPan}.
Fig.~4.32a gives a comparison \cite{HwaPan} of $\phi_q$ (crosses) and
$(q-1-\tau_q^{{\mathrm{dyn}}})$ (full circles) from ECCO simulation results and shows
that the deviation of $\tau_q^{{\mathrm{dyn}}}$ from $q-1$ is indeed close to the
deviation
of $\phi_q$ from zero. This observation gains support from the UA1 \cite{UA1G},
hA \cite{Ghosh92} (Fig.~4.32b) and AA \cite{Seng90} analysis.
The remaining difference can be attributed to the difference in the
definition of $\langle F_q\rangle$ and $\langle G_q\rangle$.
Fig.~4.32a, however, also shows that $\tau_q^{{\mathrm{dyn}}}$ cannot be simply
replaced by $\tau_q$ (open circles) in a quantitative analysis. This
is in agreement with the observation of Subsect.~4.7.5, but has not been
taken into proper consideration in
recent experimental application on lh \cite{EMCG},
hA \cite{Shiv93} and AA \cite{Sark93-2} collisions.
To summarize the present experimental findings, the data indicate that the
multifractal spectral function $f(\alpha)$ has, at least for large bin sizes,
the properties expected from the theory of multifractals. The function
$f(\alpha)$ is very sensitive to violations of universality and to details of
present Monte-Carlo models. However, with the methods used so far, the
multifractality analysis breaks down at finer resolution. The finite
multiplicity effect--- statistical noise---overwhelms and it is difficult
to disentangle it from dynamical features. An advantage is that the
$\langle G_q\rangle$ can probe holes in the distribution, not just spikes.
A recent extension of the definition of the $G$-moments filtering out
high multiplicities can claim some success in extracting the dynamical
component. The advantage is that $\langle G_q\rangle^{{\mathrm{dyn}}}$ lends itself more
readily to (multi)fractal interpretation and direct extraction of the
R\'enyi dimensions according to (2.114), while $\langle F_q\rangle$ is more
closely related to the correlation function. Fig.~4.32 can serve as a rough
link between the two. Whereas a higher-dimensional analysis could be
useful also here, the low average multiplicity even in the highest
energy experiments presently precludes further progress in this direction.
\subsection{Universal multifractals}
At first sight interesting approaches, recently
applied~\cite{Ratti92} to obtain the degree of
multifractality (or L\'evy-index) $\mu$ in multiparticle production, are
the methods of
Probability Distribution Multiple Scaling (PDMS) and Double Trace Moments
(DTM)~\cite{Ratti91}.
In the first method,
the fundamental scaling law is written in
terms of the probability for the number of particles $n_m$
in bin $m$ at resolution $M$ to be larger than a certain threshold $
n_{{\mathrm{th}}}=M^\gamma$,
\begin{equation}
P(n_m(M)> M^\gamma)\propto M^{-c(\gamma)}\ .\label{large:dev}
\end{equation}
The statistical function $c(\gamma)$ is the codimension function describing the
sparseness of large intensities $n_m$. Like the factorial moments
(\ref{dr:44}-\ref{dr:46})
or the extended $G$-moments (\ref{ggmodif}), the DPMS method is
a straightforward filter for spikes of large $n_m$.
The PDMS method is closely related to Large Deviation Theory, a topic
in probability theory and of much theoretical interest in statistical
mechanics~\cite{Ellis:l}. Equation (\ref{large:dev}) expresses a
Level-2 Large Deviation property, describing deviations of the
``empirical measure'' $n_m$ from the infinite sample probability
density; $c(\gamma)$ is related to a generalized entropy.
Fig.~4.33a shows data~\cite{Ratti92} at different charge multiplicity
$(n=6$ and 14 are given as examples) presented in a double-logarithmic plot,
for various threshold values $n_{{\mathrm{th}}}=M^\gamma$. In spite of limitations on
statistics and on multiplicity $n_m$, a region of linearity
can be seen for all multiplicities and thresholds.
This is claimed to be evidence for PDMS.
When $c(\gamma)$ is smaller than the topological dimension $D$ of the embedding
space, it is possible to define a function $D(\gamma)=D-c(\gamma)$ corresponding to
the classical fractal dimension. If a sample of $N_\ifmath{{\mathrm{s}}}$ events is used in the
analysis (instead of one event), $c(\gamma)$ can become larger than the
topological dimension since different events can contribute to the same
bin $m$. In Fig.~4.33b the function $c(\gamma)$ is shown for the same multiplies.
The dotted line corresponds to $c(\gamma_\ifmath{{\mathrm{s}}})=D+D_\ifmath{{\mathrm{s}}}$, where $D_\ifmath{{\mathrm{s}}}$ is the
sample dimension defined as $N_\ifmath{{\mathrm{s}}}=M^{D_\ifmath{{\mathrm{s}}}}$. For $n=20$, e.g., this limit
is crossed for a threshold of $n_{{\mathrm{th}}}$=3 with $N_\ifmath{{\mathrm{s}}}=15$. Singularities
of that type are called ``wild" singularities (not arising from
Poisson-like fluctuations).
A parametrization of $c(\gamma)$ in terms of two parameters is provided
by the theory of universal multifractals~\cite{Ratti91},
\begin{eqnarray}
c(\gamma) & = & C_1\left({\gamma\over \mu'C_1}+{1\over\mu}\right)^{\mu'}
{\rm \ for\ } \mu\not=1\nonumber \\
& = & C_1 \exp \left({\gamma\over C_1}-1\right)
{\rm \ for\ } \mu=1\ ,
\end{eqnarray}
with ${1\over\mu}+{1\over \mu'}=1$ and $0\leq\mu\leq2$; $\mu$ is the
L\'evy-index giving the degree of multifractality and $C_1$ is
the codimension of the average field. The two parameters
can be obtained from fits to the results given in Fig.~4.33a,
but turn out to be highly correlated.
The L\'evy-index $\mu$ can, however, be determined independently of $C_1$
with the help of the Double Trace Moments~\cite{Ratti91}, a generalization
of the $G$-moments. They are defined in~(\ref{tr}). A DTM analysis of
$\sqrt{s}=16.7$ GeV data~\cite{Ratti92} yields $\mu$-values ranging from
0.4 to about 0.9, increasing with multiplicity $n$. Such
values are far from monofractality $(\mu=0)$, but considerably below
$\mu\sim1.6$ obtained in Subsect.~4.3.3 from factorial moments.
For multifractal theory, it is important to know whether the limit
$\mu=1$ is crossed (signalling ``hard unbounded'' singularities) or
asymptotically approached from below (indicating
``soft bounded'' singularities).
This question cannot be answered at low energies and
needs high-energy, high-multiplicity data.
Extension to higher-dimensional space, though difficult in practice,
is necessary since singularities can easily be washed out if a particular
variable is not sensitive to them.
In spite of the interesting potential of the ``Universal Multifractal''
idea, one should keep in mind that the method suffers from
the same limitations as the multifractal method based on $G$-moments.
This is easily illustrated by considering again the Bernoulli-trials
model discussed in Subsect.~4.7.5, above. For the pure binomial noise
(4.34), one has
\begin{equation}
P_B(n_m(M)\geq n_{{\mathrm{th}}})=I_{1/M}(n_{{\mathrm{th}}},n-n_{{\mathrm{th}}}+1),\label{mf:1}
\end{equation}
where $I_x(a,b)$ is the incomplete beta function~\cite{abramowitz}.
\noindent The logarithm of (\ref{mf:1}) is approximately linear
in $\ln M$ for reasonably large $M$. Equation (\ref{mf:1}) not only has
all the features of the data plotted in Fig.~4.33a, but even agrees
numerically quite well. The co-dimension function $c(\gamma)$ is, for
this simple model, approximately equal to $n_{{\mathrm{th}}}$, implying constant
differences between the slopes for successive values of $n_{{\mathrm{th}}}$. The
data in Fig.~4.33b show exactly this property.
Double Trace Moments are easily calculated in the Bernoulli model. We find
that the ``L\'evy-index'' $\mu$ is a smoothly increasing function of the
event multiplicity $n$ crossing the ``hard unbounded'' value $\mu=1$ near
$n=30$. We conclude that the data in Fig.~4.33 are merely reflecting
statistical noise.
Dynamically useful information could possibly be extracted if
Double Trace {\it factorial moments} were used instead of
the usual moments. This is easily verified on the simple
Bernoulli model and, of course, applies to $G$-moments as well.
\section{Density and correlation strip-integrals}
\subsection{The method}
A fruitful recent development in the study of density fluctuations is the
density and correlation strip-integral method~\cite{Hen83,Drem88,Lipa91-53}.
By means of integrals of the inclusive density over a strip domain, rather
than a sum of box domains, one not only avoids unwanted side-effects, such
as splitting up of density spikes, but also drastically increases the
integration volume (and therefore the accuracy) at a given resolution.
Consider first the (vertical) factorial moments $F_q$ defined,
for an analysis in one dimension, as
\begin{equation}
F_q (\delta y)\equiv\frac{1}{M} \sum^M_{m=1} \frac{\langle n^{[q]}_m\rangle}
{\langle n_m\rangle^q}
= \frac{1}{M} \sum^M_{m=1} \frac{\displaystyle \int_{\Omega_m}\Pi_i dy_i\rho_q(y_1\dots y_q)}
{\displaystyle \int_{\Omega_m}\Pi_i dy_i \rho_1(y_1)\dots \rho_1(y_q)}\ \ . \end{equation}
\noindent
The integration domain $\Omega_{\PB}=\sum^M_{m=1}\Omega_m$ thus consists of $M$
$q$-dimensional boxes $\Omega_m$ of edge length $\delta y$. For the case
$q=2$, $\Omega_{\PB}$ is the domain in Fig.~4.34a.
A point in the $m$-th box corresponds to a pair $(y_1,y_2)$ of
distance $|y_1-y_2|<\delta y$ and both particles in the same bin $m$.
Points with $|y_1-y_2|<\delta y$ which happen {\it not} to lie in the same
but in adjacent bins (e.g. the asterix in Fig.~4.34a) are left out.
The statistics can be approximately doubled by a change
of the integration volume $\Omega_{\PB}$ to the strip-domain of Fig.~4.34b.
For $q>2$, the increase of
integration volume (and reduction of squared statistical error)
is in fact roughly proportional to the order of the correlation.
The gain is even larger when working in two or three phase-space
variables.
In terms of the strips (or hyper-tubes for $q>2$), we define as (vertical)
{\it density} integrals
\begin{equation}
F^\ifmath{{\mathrm{S}}}_q (\delta y) \equiv \frac{\displaystyle \int_{\Omega_{\ifmath{{\mathrm{s}}}}}\Pi_i \ifmath{{\mathrm{d}}} y_i
\rho_q(y_1,\dots,y_q)}
{\displaystyle \int_{\Omega_{\ifmath{{\mathrm{s}}}}}\Pi_i \ifmath{{\mathrm{d}}} y_i \rho_1(y_1)\dots \rho_1(y_q)}\ \
\label{3.45}
\end{equation}
and, similarly, the {\it correlation} integrals $K^\ifmath{{\mathrm{S}}}_q(\delta y)$ by replacing
the density $\rho_q(y_1,\dots,y_q)$ by the correlation function
$C_q(y_1,\dots,y_q)$. (Note that in the literature the term ``correlation
integral'' is often also used for the $F^\ifmath{{\mathrm{S}}}_q(\delta y)$.)
These integrals can be evaluated directly from the data, after selection
of a proper distance measure $(|y_i-y_j|,[(y_i-y_j)^2+(\phi_i-\phi_j)^2]^{1/2}$,
or better the four-momentum difference $Q^2_{ij} = -(p_i-p_j)^2$) and after
definition of a proper multiparticle topology, the snake integral
{}~\cite{CaSa89}, the GHP integral~\cite{Hen83}, or the star integral
\cite{Egger93} as shown in Figs.~4.34c-e, respectively.
\subsection{Results}
As an example, $F_4(\delta y)$ is compared to $F^\ifmath{{\mathrm{S}}}_4(\delta y)$ (and
$F^\ifmath{{\mathrm{S}}}_2,F^\ifmath{{\mathrm{S}}}_3)$)
for the NA22 spike event~\cite{AdamPL185-87} in Fig.~4.35a (no error-bars are
shown, because it is one event). Depending on whether the prominent spike lies
entirely in one bin or is split across two, $F_4$ shows large fluctuations.
These are practically absent in $F^\ifmath{{\mathrm{S}}}_4$ (Fig.~4.35b). Large improvement in
one-dimensional $(\eta)$ and two-dimensional $(\eta-\varphi)$ analysis is also
observed in \cite{Jie95}.
How much the statistical errors are reduced can be seen on Fig.~4.36a
where the NA22 data~\cite{Ajin89-90} are plotted as a function of
$-\ln Q^2$, with all two-particle combinations in an $n$-tuple
having $Q^2_{ij}<Q^2$ ~\cite{Hen83}. The following
observations can be made:
i) the errors and fluctuations are indeed largely reduced, as compared e.g. to
Fig.~4.20a.
ii) with the (one-dimensional) distance measure $Q^2$, the moments show a
similarly steep rise as in the three-dimensional analysis (e.g. Fig.~4.11c).
iii) Contrary to the results in rapidity, positives and negatives behave very
similarly here (only negatives are shown in Fig.~4.36a), but are now much
steeper than all-charged.
iv) $F^\ifmath{{\mathrm{S}}}_2$ is flatter for $(+-)$ than for all-charged or like-charged
combinations.
The first two observations demonstrate the strength of the new method
and the advantage of using the proper variable. The second two observations
directly demonstrate the large influence of identical particle correlations
on the factorial moments. These results agree very well with results from
the UA1 collaboration ~\cite{Alba90} shown in Fig.~4.36b and with lh results
\cite{Arne86,Adams94}.
Monte-Carlo simulations with FRITIOF~2 show the following (see Fig.~4.37
for the case of $F^\ifmath{{\mathrm{S}}}_2$). The default ``plain'' version is unable to
describe the all-charged NA22 data, but a ``biased'' version (including
misidentified Dalitz decay + 0.25\% undetected $\gamma$-conversions) comes
closer to the data. However, not unexpectedly, both versions fail completely
in describing the like-sign data, where the model stays way too low. On the
other hand, $F^\ifmath{{\mathrm{S}}}_2$ for the $(+ -)$ combination is largely overestimated
when $\gamma$-conversions are included, but saturates without.
\subsection{Transverse-momentum and multiplicity dependence}
As in Subsect.~4.4.2 (Fig.~4.20c), UA1 \cite{Wu94} has studied the $\phi_2$
dependence on the average transverse momentum $\bar p_\ifmath{{\mathrm{T}}}$ of the event,
but now in terms of density integrals in $Q^2$. In contrast to the strong
decrease (and subsequent slight increase of $\phi_2$) with increasing
$\bar p_\ifmath{{\mathrm{T}}}$ observed for the one-dimensional analysis in Fig.~4.20c,
a strikingly flat behaviour (and slight increase above 0.6 GeV/$c$) is
observed for the data (full circles) in Fig.~4.38a.
The discrepancy of PYTHIA (open circles) is even stronger here than in
Fig.~4.20c. The slope $\phi_2$ starts at even negative values for small
$\bar p_\ifmath{{\mathrm{T}}}$, but increases fast with increasing $\bar p_\ifmath{{\mathrm{T}}}$ to reach values
overestimating $\phi_2$ at $\bar p_\ifmath{{\mathrm{T}}}\ ^>\hs-2.5mm_\sim\ 0.5$ GeV/$c$.
A similar disagreement is observed for the multiplicity dependence in
Fig.~4.38b. While the UA1 data (full circles) decrease with increasing $n$,
PYTHIA predicts a strong increase. In Figs.~4.38~c) and d), it is shown that
this violent discrepancy between PYTHIA and data is mainly due to like-sign
pairs, so to the way Bose-Einstein correlations are incorporated
into the model.
\subsection{Bose-Einstein correlations versus QCD effects}
Of particular interest is a comparison of hadron-hadron to $\re^+\re^-$ results in
terms of same and opposite charges. This is shown in Fig.~4.40 for
$q=2$ UA1 and DELPHI data in~\cite{Mandl92} (note that in this figure the
derivative of (\ref{3.45}) is presented in small $Q^2$ bins). An important
difference between
UA1 and DELPHI can be observed on both sub-figures: For ``large''
$Q^2(>0.03$ GeV$^2$), where Bose-Einstein effects do not play a role,
the $\re^+\re^-$ data increase much faster with increasing $_2\log(1/Q^2)$ than
the hadron-hadron results. For $\re^+\re^-$, the increase in this $Q^2$ region is
very similar for same and for opposite sign charges. At small $Q^2$,
however, the $\re^+\re^-$ results approach the hadron-hadron results. The authors
conclude that for $\re^+\re^-$ at least two processes are responsible for the
power-law behaviour: Bose-Einstein correlations at small $Q^2$ following
the evolution of jets at large $Q^2$. In hadron-hadron collisions at
present collider energies only Bose-Einstein effects seem relevant.
Since string fragmentation causes an anti-correlation between same-charged
particles, it is of interest to compare $\re^+\re^-$ results to JETSET in terms of
strip integrals for the different charge combinations, separately.
This has been done in~\cite{Mandl92} and, indeed, the Monte-Carlo
results level off at small $Q^2$ and fall below the data for the same-charge
results, while they describe the opposite-charge data perfectly well
(not shown here).
The exact functional form of $F^\ifmath{{\mathrm{S}}}_2$ is derived from the
data of UA1~\cite{Alba90} and NA22~\cite{Ajin89-90},
again in its differential form\footnote{In fact in this differential
form $F^\ifmath{{\mathrm{S}}}_2(Q^2)$ is identical to $R(Q^2)$ usually used in Bose-Einstein
analysis. The only difference is that it is plotted on a double-logarithmic
plot, here.}, in Fig.4.40. Clearly, the data favour a power law in $Q$
over an exponential, double-exponential or Gaussian law.
If the observed effect is real, it supports a view recently developed
in~\cite{Bial92}. There, intermittency is explained from Bose-Einstein
correlations between (like-sign) pions. As such, Bose-Einstein correlations
from a static source are not power behaved. A power law is obtained i) if
the size of the interaction region is allowed to fluctuate, and/or ii) if
the interaction region itself is assumed to be a self-similar object
extending over a large volume. Condition ii) would be realized if parton
avalanches were to arrange themselves into self-organized critical
states~\cite{Bak87}. Though quite speculative at this moment, it is an
interesting new idea with possibly far-reaching implications.
We should mention also that in such a scheme intermittency is viewed as
a final-state interaction effect and is, therefore, not troubled by
hadronization effects.
The effect on the factorial moments of adding Bose-Einstein correlations in
FRITIOF is convincingly demonstrated for heavy-ion collisions
in~\cite{Kadi92}. Because of the large number of collision processes, other
correlation effects are expected to play a much reduced role for this type
of interaction and Bose-Einstein correlations, as a collective effect, can
become the dominant source of non-statistical fluctuations. Also from these
results it is clear that more than one fixed interaction-volume
radius is needed to reproduce the experimental results.
In perturbative QCD, on the other hand, the intermittency indices $\phi_q$,
are directly related to the anomalous multiplicity dimension
$\gamma_0=(6\alpha_\ifmath{{\mathrm{s}}}/\pi)^{1/2}$ \cite{Gust91,Oc92,OW92a,dd92,BrMeuPe93} and,
therefore, to the running coupling constant $\alpha_\ifmath{{\mathrm{s}}}$. In the same theoretical
context, it has been argued \cite{Oc92,OW92a,dd92,BrMeuPe93} that the
opening angle $\chi$ between particles is a suitable and sensitive variable
to analyse and well suited for these first analytical QCD calculations of
higher-order correlations. It is, of course, closely related to $Q^2$.
A first analytical QCD calculation \cite{Oc92,OW92a} is based on the
so-called double-log-approximation with angular ordening \cite{ao} and
on local parton-hadron-duality \cite{lphd}. A preliminary comparison with
DELPHI data \cite{ManBu94} gives encouraging results, even including an
estimate for the running of the strong coupling constant $\alpha_\ifmath{{\mathrm{s}}}$.
\section{Correlations in invariant mass}
\defM_{\mbox{\small inv}}{M_{\mbox{\small inv}}}
\defK_2^{+-}(\MINV){K_2^{+-}(M_{\mbox{\small inv}})}
\defK_2^{--}(\MINV){K_2^{--}(M_{\mbox{\small inv}})}
The previous section has illustrated the advantages of the correlation
integral method with a ``distance'' measure directly related to
the invariant mass of the particle system. The results give additional
support to the {Fia\l kowski} conjecture, mentioned in Subsect.~4.3.2,
from which could be anticipated that dynamical effects are most clearly
revealed if the correlation functions and factorial moments are
directly analysed in terms of Lorentz-invariant variables.
Evidently, there are many arguments in favour of invariant mass as a
dynamical variable rather than the single-particle variables
often used in early intermittency studies. Resonances, the cause of
most of the correlations among hadrons, and threshold effects
appear at fixed values of mass; Bose-Einstein interference correlations
depend on four-momentum differences; multiperipheral-type ladder
diagrams are functions of two-particle invariant masses, and so on.
The idea to study correlations as a function of invariant mass was,
to our knowledge, first proposed in~\cite{berger,thomas}.
The authors introduce a method which is technically a differential
version of the correlation integral method. It focusses directly on
the correlation functions (cumulants) rather than on the inclusive
density as in (\ref{3.45}). Starting from the definition (\ref{dr:13}),
one defines the correlation function
\begin{equation}
C_2(M_{\mbox{\small inv}})=\rho_2(M_{\mbox{\small inv}})- \rho_1\otimes\rho_1(M_{\mbox{\small inv}}),\label{CM}
\end{equation}
obtained after integration (in a suitable region of phase space)
of $C_2(p_1,p_2)$ over all variables except $M_{\mbox{\small inv}}$. Here,
$\rho_2(M_{\mbox{\small inv}})$ is the familiar normalized
2-particle invariant-mass spectrum. The ``background term''
$\rho_1\otimes\rho_1(M_{\mbox{\small inv}})$ is the integral
of $\rho_1(p_1)\rho_1(p_2)$ with $M_{\mbox{\small inv}}$ fixed.
For the data shown below, it is obtained from ``uncorrelated'' (``mixed'')
events, built by random selection from a track pool. The same method is
used in evaluating the denominator in (\ref{3.45}).
Higher-order correlations are obtained in a completely analogous manner.
We further utilize the function
$K_2(M_{\mbox{\small inv}})=C_2(M_{\mbox{\small inv}})/\rho_1\otimes\rho_1(M_{\mbox{\small inv}})$, the normalized
factorial cumulant of order two.
The analysis in \cite{berger}, based on low statistics pp data at 205
GeV/$c$, demonstrates that $K_2^{+-}(M_{\mbox{\small inv}})$ and $K_2^{\pm\pm}(M_{\mbox{\small inv}})$
follow an approximate power law, written by the authors as
\begin{equation}
K_2(M_{\mbox{\small inv}})=(M_{\mbox{\small inv}}^2)^{\alpha_\ifmath{{\mathrm{X}}}(0)-1}\label{power-law}.
\end{equation}
The notation reminds of the interpretation of (\ref{power-law})
in terms of the Mueller-Regge formalism (for details see~\cite{berger}).
The power $\alpha_\ifmath{{\mathrm{X}}}(0)$ is the appropriate Regge-intercept, $\ifmath{{\mathrm{X}}}=\ifmath{{\mathrm{R}}}$
for non-exotic pairs and $\ifmath{{\mathrm{X}}}=\ifmath{{\mathrm{E}}}$ for exotic ones.
The ratio $K_2^{--}/K_2^{+-}$ was further seen to fall as
$M_{\mbox{\small inv}}^{-2}$, consistent with $\alpha_\ifmath{{\mathrm{R}}}(0)-\alpha_\ifmath{{\mathrm{E}}}(0)=1$.
Not relying on Mueller-Regge theory, the authors
argued that most of the correlations at small $M_{\mbox{\small inv}}$
are due to resonance decays into three or more pions and
to interference of amplitudes~\cite{thomas}.
The results already obtained in~\cite{berger} clarify several issues which
have troubled the interpretation of intermittency data. Among others, they
demonstrate that different charge-states should be treated separately since
the $M_{\mbox{\small inv}}$ dependence is very different. This fact, obvious in $M_{\mbox{\small inv}}$
but much less so in rapidity, was not fully appreciated in early
intermittency analysis and the crucial importance of like-sign particle
correlations remained hidden in ``all-charged'' analyses.
The method of~\cite{berger} has now been applied by NA22 \cite{XX1}
and DELPHI \cite{XX2}. Figure~4.41 shows data
on $K_2(M_{\mbox{\small inv}})$ for a combined sample of non-diffractive $\pi^+/\PK^+p$
collisions at 250 GeV/$c$ in the central c.m.~rapidity region $-2<y<2$.
$K_2^{+-}(\MINV)$ has a prominent $\rho^0$ peak, but is quite flat near threshold.
The peak in the first bin of sub-figure~(a) is attributed to
contamination from Dalitz-decays and $\gamma$-conversions.
$K_2^{--}(\MINV)$ falls much faster. A fit of $K_2\sim (M_{\mbox{\small inv}}^2)^{-\beta}$ yields
$\beta^{--}=1.29\pm0.04$, $\beta^{++}=1.46\pm0.03$, $\beta^{+-}=0.17\pm0.02$,
in agreement with~\cite{berger} and consistent with the relation
$\alpha_R(0)-\alpha_E(0)=1$.
NA22 also finds that cuts on transverse momentum or relative azimuthal angle
$\delta\varphi$ strongly affect the shape of $K_2^{+-}(\MINV)$, but have little effect
on $K_2^{--}(\MINV)$ for $M_{\mbox{\small inv}}<0.5$ GeV/$c$${}^2$. This means that $K_2^{--}(\MINV)$ at small
$M_{\mbox{\small inv}}$ is essentially a function of $M_{\mbox{\small inv}}$ (or $Q^2$) only, illustrating
once more the advantage of $M_{\mbox{\small inv}}$ compared to other variables.
The data in Fig.~4.41 confirm the conclusion of Sect.~4.8 that
the correlations in like-charge systems are at the origin of
the strong increase of factorial moments for small invariant masses.
Whether Bose-Einstein effects are solely responsible for the
differences between ($\pi^{\pm}\pi^{\pm}$) and $(\pi^+\pi^-)$ pairs
is not so evident. It suffices to consider~\cite{thomas}
the contributions from decays of various resonances to realize that
the $M_{\mbox{\small inv}}$-dependence near threshold for ``exotic'' particle systems
must be stronger than for ``non-exotic'' ones. In a dual Regge picture,
such differences translate into very different values of the
respective Regge intercepts as in (\ref{power-law}).
It remains, therefore, to be verified if the $M_{\mbox{\small inv}}$-dependence of
the data can be explained as a superposition of a ``standard'' Regge-type
power law and a conventional Bose-Einstein enhancement.
As pointed out in \cite{XX3}, there is a feasible way to test this and
even to give access to the relative strength of BE interference and exotic
like-charge $\pi\p$ interaction. The idea is that particle combinations exist
which are either a) exotic, but not identical (e.g., $\PK^+\pi^+$ or
$\PK^-\pi^-$ pairs) or b) identical, but not exotic ($I=0$ $\pi^0\pi^0$ pairs).
NA22 \cite{XX4} and ALEPH \cite{XX5} data indicate that very-short-range
correlations are indeed absent in the exotic $\PK\pi$ channel. This supports
Bose-Einstein correlations rather than exotic Regge behaviour, but the point
deserves further investigation.
The possibility that the correlation functions depend mainly on
invariant mass has interesting further consequences. These were
analysed in~\cite{edw:sant}. Taking in (\ref{CM})
\begin{equation}
C_2(M_{\mbox{\small inv}})\propto (M_{\mbox{\small inv}}^2)^{\alpha-1}\,\mbox{BE}(M_{\mbox{\small inv}}),\label{edw:c2}
\end{equation}
with $\mbox{BE}$ a conventional Bose-Einstein factor, exponential in $Q$,
good agreement is obtained with the NA22 second-order correlation integral
data of ($--$)-pairs. Integrating the correlation function over all
variables except $\delta y$ gives $F^{--}_2(\delta y)$ which also fits the
data. Although $C_2$ does not explicitly depend on the transverse momentum
of the particles, it turns out that $F_2(\delta y,p_{\ifmath{{\mathrm{T}}} 1},p_{\ifmath{{\mathrm{T}}} 2})$ is
larger and more steeply increasing than $F_2(\delta y)$ for small
$\delta y$ and small $p_\ifmath{{\mathrm{T}}}$'s. The opposite happens for large $p_\ifmath{{\mathrm{T}}}$'s.
This is the ``low-$p_\ifmath{{\mathrm{T}}}$ intermittency effect'' seen in the NA22 and UA1
data (cfr.~Fig~4.20a and Subsect.~4.4.2). The explanation is simple: under
the stated hypothesis, small $p_\ifmath{{\mathrm{T}}}$ for the two particles in a pair
means, on the average, smaller invariant mass than for unrestricted
transverse momentum and, therefore, larger and shorter-ranged correlations
in rapidity. Enhanced intermittency follows as a consequence of kinematical
cuts! The influence of the Bose-Einstein factor is easily checked in this
simple model. It is found to be necessary in order to reproduce the
correlation integral data and $F_2$ for restricted $p_\ifmath{{\mathrm{T}}}$ but has, as
expected a priori, very little influence on the $p_\ifmath{{\mathrm{T}}}$-integrated
$F_2(\delta y)$. This explains early controversy over the role of
Bose-Einstein effects in one-dimensional factorial moment analyses
(Subsect.~4.4.1).
A study of the invariant-mass dependence of the two-particle correlation
function has for the first time given clear indications as to why the
hadron-hadron Monte-Carlo models fare so badly when confronted with
factorial moment data. For example, Fig~4.42 shows NA22 data for
$K_2^{--}(M_{\mbox{\small inv}})$ and $K_2^{+-}(M_{\mbox{\small inv}})$ compared to FRITIOF. The
predicted shape of $K_2^{+-}(M_{\mbox{\small inv}})$ is very different from the data,
especially in the $\rho^0$ region. It shows an enhancement at low mass
which causes the correlation function to drop much faster than seen in
the experiment. In the model, this structure is traced back to reflections
from $\eta$, $\eta'$ and $\omega$ resonances. The model also fails to
describe $K_2^{--}(M_{\mbox{\small inv}})$ since correlations are very weak or even
negative, except for a threshold enhancement due to $\eta'$-decays.
These examples suffice to demonstrate that FRITIOF (or rather JETSET)
has serious shortcomings and is unable to reproduce two-particle correlations
in invariant mass. For correlations in rapidity and azimuthal angle
this was seen earlier (Sect.~3.1), but the reasons remained obscure,
mainly because of the insensitivity of these variables to dynamical
correlations at small mass.
A study of the correlation function in terms of invariant mass clarifies
the situation considerably. For NA22, the model was known to
overestimate significantly the production rates of
$\rho^0$ and $\eta$ mesons~\cite{na22:atayan} and presumably also
those of $\eta'$ and $\omega$ for which no direct measurements
exist (see also~\cite{Walker91} for h$A$ collisions).
This is now seen to distort heavily the $M_{\mbox{\small inv}}$-dependence of $K_2$.
Also Bose-Einstein low-mass enhancements, most likely responsible
for the fast drop of $K_2^{--}(\MINV)$ in the threshold region, are not
included in the FRITIOF model commonly used. Finally, we note
that the values of $K_2$ in the considered mass interval are much smaller
than the data. This is related to the width of the charged particle
multiplicity distribution which is known to be too small in FRITIOF.
It affects the global magnitude of factorial moments and cumulants.
In Fig.~4.43a, the discrepancy is shown to be quite similar for $(+-)$
correlations in $\re^+\re^-$ collision \cite{XX2} and JETSET. As in hh collisions,
the correlation is underestimated in the mass region below the $\rho^0$. This
discrepancy can be cured by decreasing the $\eta'$ and $\rho^0$ production and
increasing $\omega$ production in JETSET.
Fig.~4.43b gives the like-sign correlation for the data and JETSET without
BE correlation. For $M_{{\mathrm{inv}}}<0.6$ GeV/$c^2$. The experimental data are
considerably higher than JETSET. This can be attributed to Bose-Einstein
interference. However, it is striking that JETSET also predicts a strong
rise towards threshold even without Bose-Einstein correlations. This is
the tail of the QCD effect also seen in Fig.~4.39 and mainly due to
multijet events. The difference between JETSET and data can indeed be
removed by including BE correlations in the model, but the $Q^2$ cut used
$(Q^2>0.04^2$) is too high to be able to distinguish a power law from
an exponential or Gaussian, as is done in Fig.~4.40.
To summarize, the above proves that the failures of models such as
FRITIOF and JETSET with respect to factorial moment and correlator data
(Subsects.~4.2.3 and 4.6.3), are not necessarily due to ``novel'' dynamics.
They are in first instance a consequence of a variety of
defects---such as incorrect resonance production rates and
absence of identical particle symmetrization---which belong
to ``standard'' hadronization phenomenology.
These defects should be eliminated before ``new physics'' can be claimed.
For $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ annihilation at LEP energies, we have found that
models such as JETSET-PS are much more successful than for all other
processes. Besides the evident fact that this process is much better
understood theoretically, QCD effects dominate and the model
parameters are much better tuned to the data. Still, serious, recently
observed discrepancies of e.g.~JETSET-PS with LEP measurements on
particle and resonance production rates are a clear sign that
hadronization in $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ is in fact less well understood than
commonly stated and needs improvement. It could be rewarding to
investigate carefully and differentially the invariant mass dependence
of the correlations using the sensitive methods now available.
Originating mainly from the low invariant mass region (typically
$<1.5$ GeV/$c$${}^2$), it is not impossible that the observed correlations
are quite independent of the process initiating the primary colour separation
in the collision, being dominated by strong final-state interactions.
This would explain ``universality'' in the sense discussed earlier.
Many authors argue that ``intermittency'' is somehow connected to (nearly
scale-invariant) perturbative QCD-cascading. Others strongly contest this
view on the argument that QCD cascades have a limited extent even at LEP
energies and are dominated by a very small number of ``hard'' emissions.
In the former case, one may expect significant differences in the
correlation functions at low mass for $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$, on the one hand, and
for hh, h$A$ and $AA$ collisions, on the other. Preliminary
$\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ data, mentioned in Sect.~4.8, seem to support the last opinion.
Whatever the final outcome, if differences are found, they should be used to
clarify the respective roles of perturbative and hadronization phases
in the different types of collision processes.
\section{Genuine higher-order correlations}
Multiparticle production in high-energy collisions is one of the rare
fields of physics where higher-order correlations are directly accessible
in their full multi-dimensional characteristics, under well controlled
experimental conditions.
Three-particle correlations have been observed in the form of short-range
rapidity correlations and higher-order Bose-Einstein correlations, but
evidence for {\it genuine} higher-order correlations (i.e., after
subtraction of all lower-order contributions) is very limited.
While it was found to be completely absent in heavy-ion collisions, first
evidence was given in Sect.~4.5 for their existence in hh collisions.
The correlation integral method turns out particularly useful for the
unambiguous establishment of genuine higher order correlations in terms
of the normalized cumulants $K_q(Q^2)$, when using the star integration
\cite{Egger93}
\begin{equation}
K^*_q(Q^2) = \frac{\displaystyle \int \Pi_i dy_i \Theta_{12}\dots\Theta_{1q}C_q(y_1,\dots,y_2)}
{\displaystyle \int \Pi_i dy_i \Theta_{12}\dots\Theta_{1q} \rho_1 (y_1) \dots \rho_1(y_q)}\ \ ,
\end{equation}
with $\Theta_{1j} = \Theta(Q^2-Q^2_{ij})$ restricting all $q-1$ distances
$Q^2_{1j}$ to lie within a distance $Q^2$ of the position of particle 1.
The star-integral method combines the advantage of optimal use of available
statistics and minimal use of computer time. Since higher acccuracy is
abtainable, dynamical structures in the correlations can be studied in
greater detail than with conventional methods.
Non-zero values of $K^*_q(Q^2)$ increasing according to a power law with
decreasing $Q^2$ are indeed observed for E665 for third order \cite{Adams94}
and for NA22 up fifth order \cite{genuine} (see Fig.~4.44 for the latter).
\section{Summary and conclusions}
\begin{enumerate}
\item Intermittency, defined as an increase of normalized factorial moments
with increasing resolution in phase space, is seen in all types of
collision. Intermittency is a 3D phenomenon. The anomalous dimensions
are small ($d_q$=0.01-0.1) in a one-dimensional analysis, but the
factorial moments are considerably larger, and their
resolution-dependence more power-like, in two- or three-dimensional
phase space. Self-similarity in the dynamics of multiparticle
production is an attractive but not fully proven explanation.
\item The factorial-moment method is very sensitive
to biases in the data. These have to be
studied in detail before final conclusions can be drawn.
Because of its sensitivity, the method has in fact proven to be
very helpful in detecting and tracing such biases.
\item The logarithms of factorial moments satisfy a possibly
dimension-independent linear relation which allows to determine
directly ratios of anomalous dimensions. The observed order-dependence
of anomalous dimensions excludes a second-order phase transition
(as treated in~\cite{BialHwa91}) as the origin of intermittency. Valid
in random cascade models, the Ochs-Wo\v siek relation shows that
correlation functions of different order are inter-related in a specific
hierarchical structure. An explanation in terms of dynamics
has not yet been found.
\item In hadron-hadron collisions, factorial moments and intermittency
indices depend strongly on transverse momentum and are largest for
low transverse momentum hadrons. This effect, also seen in rapidity-
and azimuthal correlations, needs further study in $\re^+\re^-$ collisions
and in parton shower Monte Carlo's. There are serious indications that
``low-$p_\ifmath{{\mathrm{T}}}$ intermittency'' is a reflection of the very strong
dependence of correlation functions for identical particles on
invariant mass. This has to be examined in more detail.
\item The multiplicity dependence in hh collisions agrees with what is
expected from mixing of independent sources. However, the {Fia\l kowski}
observation on possible universality casts some doubt on this type of
interpretation. For a given density, heavy-ion collisions show more
intermittency than hadron-hadron collisions, possibly as a result of
Bose-Einstein interference or other collective effects.
\item Factorial cumulants are direct measures of genuine higher-order
correlations. These are present in hadron-hadron collisions, in particular
for small phase-space domains, but seem to be absent in heavy-ion collisions.
\item Factorial correlators reveal bin-bin correlations.
The correlators $F_{pq}$ increase with
decreasing correlation length $D$, but only approximately follow
a power law for $D\ ^<\hs-2.5mm_\sim\ 1$. For fixed $D$, the values of
$F_{pq}$ are independent of resolution $\delta y$, a property
predicted in the $\alpha$-model, but also shared by
models with short-range order such as FRITIOF. The powers
$\phi_{pq}$ increase linearly with the product $pq$ of the orders,
but are considerably larger than expected from FRITIOF and from
the simple $\alpha$-model.
\item A recent extension of the definition of $G$-moments filtering out
high multiplicities, claims success in extracting the dynamical component.
However, under the conditions prevailing in present hadroproduction
experiments, multifractal and generalized multifractal methods seem unable
to overcome the overwhelming dominance of statistical fluctuations.
\item The correlation (or density) strip integral strongly reduces
statistical errors, as well as fluctuations due to splitting of spikes.
Using the squared four-momentum $Q^2_{ij}$ as a distance measure, an
increase similar to that found in three-dimensional analyses
is observed. This increase is caused by correlations among like-charged
particles. Bose-Einstein interference must contribute significantly
to the intermittency effect but is not power-behaved in the
conventional approach. Power-law behaviour in Bose-Einstein interferometry
would imply a random superposition of ``emission centres'' with possibly
fractal properties. Parton avalanches in a self-organized critical
state are an intriguing possibility.
\item The analysis of cumulants in terms of invariant mass, or related
variables, reintroduced recently after early work, has helped in
clarifying several issues in intermittency. The reasons behind the
dramatic failures of models for hadron-hadron collisions are clearly
revealed. They are in first instance incorrectly predicted particle
and resonance production rates and the near absence of correlations
(or even presence of anti-correlations) in identical particle systems
with small invariant masses. These defects are not easy to cure in a
consistent manner by simple parameter tuning and ``new'' physics may
be needed to restore internal consistency in
e.g.~string-fragmentation models.
\item The hadronization mechanism in hadron-hadron collisions is based
on identical physical principles and Monte-Carlo algorithms as those
applied with apparently great success to ``hard'' processes, in particular
to $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ annihilation.
There is now growing experimental evidence that the mentioned discrepancies,
first seen in hadron collisions, are also present in such processes. These
should first be removed before commonly expressed claims that ``No new
physics is involved in intermittency'' can be accepted.
\item The relative importance of perturbative and non-perturbative QCD
contributions to hadron correlations in $\ifmath{{\mathrm{e}}}^+\ifmath{{\mathrm{e}}}^-$ remains controversial,
also theoretically. This can be studied further with the new techniques
now available.
\end{enumerate}
\chapter{Theoretical description}
In parallel to the extensive experimental effort in quest for power-law
behaviour, intense activity has developed on the theoretical side,
to find acceptable explanations of the rapidly accumulating collection of
data on factorial moments. The meaning of ``intermittency'' in
multiparticle processes is still the subject of much debate and no definite
consensus has as yet emerged. Let us remember once again that we are
dealing here with the problem of evolution of particle number
distributions (or multiparticle correlations) in ever smaller bins.
{\it A priori, }
the most direct road of attack starts from quantum chromodynamics (QCD),
now firmly established as the theory of strong interactions. Unfortunately,
since the problem of confinement is unsolved, QCD can only be used as a
guideline to build phenomenological models for soft hadronic phenomena.
While successful for $\re^+\re^-$-annihilation, such models remain at present
unsatisfactory for most other processes, and in particular for hh-collisions.
The model's deficiencies are often invoked in support for claims of
``new physics'', but also this matter is far from being settled.
{}From the outset it is clear that phenomena such as ``intermittency''
are manifestations of dynamics in a, most probably, strongly non-linear
regime of QCD. It is, therefore, quite likely that the observed phenomena
are not very sensitive to the precise form of the Lagrangian, even though
the general properties of the interacting fields (e.g., the vector nature
of gluons) surely play a crucial role. Hence, a satisfactory description
might be possible on the basis of quite general properties of non-linear
systems as revealed by complex systems in many other branches of physics.
This idea lies at the origin of various attempts to establish connections
with models for turbulence, multiplicative cascade processes,
``effective'' field theory, the statistical mechanics of disordered systems,
fractals, phase transitions of various kinds and others.
A perturbative QCD approach to intermittency
would provide a viable explanation if the
hadronization of quark and gluon systems possesses
the property of ``early confinement''
such that local parton-hadron duality would hold.
Various efforts in this direction have
indeed established that parton (quark-gluon)
avalanches exhibit (multi)fractal properties. These
follow from the (fortunate) fact
that the process possesses Markovian properties. Further
developments along this line will
evidently improve our understanding of perturbative
cascades.
A direct connection with experimental observations is,
however, not yet established.
Numerous other interpretations of ``intermittency''
have been advanced and will briefly be
described further below.
\section{Simplest approximations}
\subsection{The linked-pair approximation}
If the rapidity distribution does not change appreciably within a bin interval
(this is justified for small intervals, at least)
one can rewrite (\ref{dr:47}),
approximating it by
\begin{equation}
K_q(\delta y) \approx \frac{1}{M(\delta y)^q} \sum^M_{m=1} \int\limits_{\delta y}
\prod_i \ifmath{{\mathrm{d}}} y_i K_q(y_1,\dots,y_q)\ . \label{kq1}
\end{equation}
Since there are no statistically independent contributions to the cumulant
functions $C_q$ and $K_q$, their arguments should be somehow linked.
Studying the correlation of galaxies~\cite{Peeb80}, it was noted that
$K_q$ can be decomposed into sums of products of two-particle correlation
functions, $K_2$ with overlapping arguments, in such a way that
all multiparticle correlations are
expressible as successive two-particle ones, so that the
whole chain of particles becomes
correlated. The last condition is necessary since we have learned that
$q$-particle correlations ($q>2$) are indeed present in the data.
In this scheme, higher-order scaled cumulants~\cite{CaSa89} are written as
\begin{equation}
K_3(y_1,y_2,y_3) = \frac{A_3}{3} \sum_{{\mathrm{perm}}} K_2(y_1,y_2)K_2(y_2,y_3)\ ,
\label{k32}
\end{equation}
\begin{equation}
K_4(y_1,y_2,y_3,y_4) = \frac{A_4}{12} \sum_{{\mathrm{perm}}} K_2(y_1,y_2)K_2(y_2,y_3)
K_2(y_3,y_4) \ , \label{k43}
\end{equation}
etc. Here, all the permutations of indices $1,\dots,q$ are summed over;
the number of terms is equal to
the denominator of the factor in front of the sum.
The numerator is an a priori arbitrary parameter
for each order of correlation.
Even now, the numerical integration in (\ref{kq1}) is hard to perform.
For that reason it
was suggested~\cite{CaSa89} to assume translation invariance of the cumulants
$K_q$, and to use the strip approximation (i.e. instead of integrating
over a set of hyper-cubes
with linear size $\delta y$, one integrates over a strip along the main axis
$Y$ and over the differences $\zeta_i=(y_{i+1}-y_i)$). In this
way (\ref{kq1}) reduces to a simple but approximate formula~\cite{CaES91}:
\begin{equation}
K_q\approx A_q(K_2)^{q-1}\ . \label{kq4}
\end{equation}
Substitution of $K_q$ in (\ref{dr:48}) allows any factorial moment
to be expressed in terms of the
second moment so that, for example,
\begin{equation}
F_3 = 1+3K_2+A_3K^2_2\ . \label{f35}
\end{equation}
As mentioned in Sect.~4.5, one can describe ~\cite{CaES91} the UA1 and UA5
data on factorial moments at energies from $\sqrt s$=200 GeV
to $\sqrt s$=900 GeV with constant values of $A_q$ for all intervals
$\delta y$. Still, the intermittency indices derived from the linked-pair
approximation remain somewhat below the experimental ones. At the lower
energy of the NA22 experiment, one gets a much larger value of $A_3$ if
the above estimate of $F_3$ is used boldly~\cite{Ajin89-90}. However,
the assumption of translation invariance is not really justified there
and one can hardly use~(\ref{f35}) on data which are averaged over a large
region of phase space.
Apart from this problem, there are also more basic questions which remain
unanswered. For instance, even within the framework of the linked-pair
ansatz, one could add to (\ref{k32}) a term containing a product of three
$K_2$'s (loops or ring-graphs), as well as further terms with multiple
links among the pairs.
Although a dynamical justification for the linked-pair approximation
is lacking at present, it should be remembered that similar approximation
methods have proven their utility, e.g.~in the theory of liquids.
Within the Born-Bogoliubov-Green-Kirkwood-Yvon (BBGKY) hierarchy scheme,
they allow to ``close'' the otherwise infinite sequence of equations
relating correlation functions of all orders. That the linked-pair
ansatz may have more than accidental relevance is further indicated by
the non-trivial fact that (\ref{kq4}) with $A_q=(q-1)!$
corresponds to a multiplicity distribution in $\delta y$ which is
of Negative Binomial type with $k$-parameter
$k=1/K_2(\delta y)$~\cite{WolfAPP90}. This two-parameter distribution
satisfactorily describes a large variety of (non-averaged) multiplicity
data and also occurs as an approximation to the soft parton
multiplicity distribution in QCD-jets~\cite{malaza:webber,nbd:gio,LAnd}.
Further extensions of the linked-pair approach beyond
those of~\cite{CaSa89,WolfAPP90} are treated in~\cite{lvanhove90,bozek91}.
The structure of many-particle correlations has also been analysed in
partially coherent radiative systems~\cite{CaFrie89}. This
approach is closely related to the linked-pair ansatz~\cite{WolfAPP90}.
Conformal theories, treated in connection with intermittency
in~\cite{dnL92},
provide an alternative~\cite{Dremextra} to the linked-pair ansatz.
In such theories the $q$-th order irreducible Green function is
written as a product of two-particle ones to the power $1/(q-1)$.
Taking into account the $q(q-1)/2$ permutations of all particle
indices, one finds
\begin{equation}
K_q\approx B_q (K_2)^{q/2},\label{drem:extra0}
\end{equation}
instead of (\ref{kq4}), which also fits the experimental data reasonably
well~\cite{Dremextra}.
\subsection{The singularities of the correlation functions}
According to relations (\ref{dr:44})-(\ref{dr:48}),
the singular behaviour of the
factorial moments at small rapidity binning implies that the
correlation functions are singular for small separation of their arguments.
In particular,
the leading singularities of the correlation functions $\rho_2$ and $C_2$ should
coincide with the singularity of the corresponding factorial moment $F_2$, if
the formal mathematical limit $\delta y\to0$ is considered. For
\begin{equation}
F_2 \sim (\delta y)^{-\phi_2}\ \ \ (\delta y \to 0) \label{f26}
\end{equation}
one should get
\begin{equation}
C_2(y_1,y_2)\sim C^{(\ifmath{{\mathrm{L}}})}_2 (y_1,y_2)|y_1-y_2|^{-\beta}+C^{(\ifmath{{\mathrm{N}}})}_2(y_1,y_2)
\label{c27}
\end{equation}
with $\beta=\phi_2,\ \ C^{(\ifmath{{\mathrm{L}}})}_2$ is a regular function of $(y_1-y_2)$,
while $C^{(\ifmath{{\mathrm{N}}})}_2$ can contain non-leading singularities (less singular
than the first term).
A two-component model of this kind has been used in~\cite{AnCo90}, where
$C^{(\ifmath{{\mathrm{N}}})}_2$ (and similarly for the higher-order correlations) is chosen
to be a regular function which results in a constant additive term to $F_2$.
The origin of the singular term has been ascribed in~\cite{AnCo90} to a
phase transition. For a singular term to be dominant, it must overwhelm
$F_2$ even numerically. However, the experimental data discussed above
indicate that this is not the case. Linear dependence in a double-log
plot is observed over quite a large background in the available region
of $\delta y$. This implies that in this region the integral contribution
$F^{(\ifmath{{\mathrm{L}}})}_2$ of the singular terms in (\ref{c27}) to $F_2$ is rather
small so that
\begin{equation}
F^{(\ifmath{{\mathrm{L}}})}_2 \ll F_2. \label{fl8}
\end{equation}
Such a situation provides~\cite{AnCo90,Fial91} new possibilities with
$\beta$ different from $\phi_2$, since in that case one gets
\begin{equation}
\ln F_2\approx \ln (1+F^{(\ifmath{{\mathrm{N}}})}_2) + F^{(\ifmath{{\mathrm{L}}})}_2/(1+F^{(\ifmath{{\mathrm{N}}})}_2)\ .
\label{ln9}
\end{equation}
It could even suggest a logarithmic dependence of $F^{(\ifmath{{\mathrm{L}}})}_2$ on $\delta y$:
$F^{(\ifmath{{\mathrm{L}}})}_2\sim \log\delta y$, i.e. a logarithmic singularity of the correlation
function for coinciding rapidities. However, such a behaviour is
indistinguishable from a power-like one for small exponents $\beta$ and the
rather restricted range of rapidity intervals (e.g., $0.1<\delta y<1$) in which
one usually looks for this
dependence. Actually, if $\beta\log\delta y\ll 1$ one gets~\cite{Fial91}
\begin{equation}
\ln F_2 \approx \ln (1+F^{(\ifmath{{\mathrm{N}}})}_2) + \frac{\bar C^{(\ifmath{{\mathrm{L}}})}_2}{\rho^2_1}
(\delta y)^{-\beta}\approx \alpha_2-\phi_2\ln\delta y \label{ln10}
\end{equation}
where
\begin{equation}
\alpha_2 = \ln(1+F^{(\ifmath{{\mathrm{N}}})}_2) + \frac{\bar C^{(\ifmath{{\mathrm{L}}})}_2}{\rho^2_1} \ {\rm \ and\ }\
\phi_2=\beta \frac{\bar C^{(\ifmath{{\mathrm{L}}})}_2}{\rho^2_1}\ . \label{a211}
\end{equation}
Herefrom, one would expect the intermittency exponent to be much smaller
than the corresponding strength of the singularity of the correlation
function, i.e. $\phi_2 \ll \beta$.
The difference between logarithmic and power-like singularities may become
observable for higher moments at small $\delta y$ as an upward curvature
appearing on log-log plots for power-law dependence. This would be more
noticeable for smaller $\delta y$, for higher $q$ and for larger values of $\beta$.
Still, one should not enter into the region of extremely small $\delta y$, where
the empty-bin effect may dominate and turns down all the curves.
The existing experimental data on factorial moments are not in contradiction
with the above, although no clear signal of an upward curvature
of higher moments is seen because of large irregularities appearing at small
$\delta y$ and large $q$. These irregularities are suppressed if
the method of correlation integrals \cite{Hen83,Drem88,CaSa89,Alba90} is
used, where the binning procedure is
not fixed but naturally follows the event structure. This has been discussed
in Sect.~3.8. In fact, the singularities are still better exposed
if factorial cumulants (\ref{dr:47}) are used instead of factorial moments
{}~\cite{FiMPI91}.
\subsection{Intermittency and Bose-Einstein correlations}
One of the possible sources of increase of factorial moments at small bin
sizes is the well-known attraction of identical Bose-particles (pions)
when their momenta are very close. Therefore, one is tempted~\cite{Gyul90}
to the extreme supposition that "intermittency" is governed by Bose-Einstein
correlations, i.e. by symmetry properties of fields but not by their
dynamics. As was shown in Chapter~4, the experimental data do indeed give
some indications on the relevant contribution of such an effect. In general,
the introduction of BE correlations tends to reduce the disagreement between
experimental data and Monte-Carlo models. However, it was also shown there
that the dynamical part is non-negligible and, consequently,
is of main concern to us.
Referring the reader to the more specialized review
of the subject in~\cite{abds95}, we would like just to point out that the same
physics effects can become more (or less) pronounced depending on the
corresponding choice of the variable in which it is displayed. In particular,
it is shown in~\cite{abds95} that the rapidity variable is suitable to reveal
the
dynamical intermittency and to suppress the Bose-Einstein contribution to
factorial moments. At the same time, when analysed as functions of squared
4-momentum transferred between pions, the factorial moments get the power-like
increasing share of BE-correlations which shadows the true (dynamical)
intermittency. Therefore, the power-like behaviour in that variable neither
proves nor disproves dynamical intermittency. One should keep this in mind
when looking at the corresponding experimental data. Surely, for quantitative
estimates Monte-Carlo calculations are necessary with full account of
the indefiniteness in the description of the BE-effect itself.
\section{Dynamical approaches}
\subsection{Various theoretical models}
{}From a theoretical point of view, experimental results are best approached
via quantum chromodynamics (QCD). Attempts in that direction are further
described in Subsect.~5.2.6. Unfortunately, the
application of QCD to soft processes involving small momentum transfers is
quite limited since strong non-perturbative effects are involved
(unless we use additional assumptions, as the local parton-hadron
duality hypothesis generalized to correlations of any order). Hence, we
are compelled either to construct general relations like as of
Chapter~2 and the previous section, or to develop phenomenological models
that fit experimental distributions by adjusting a number of free parameters.
By now, many phenomenological models have been proposed. Ideas inspired
by QCD have been used in parton shower models and in their phenomenological
counterparts: the dual topological model and quark-gluon string
models~\cite{LUND,DPM,QGSM,CaTr88,Ande87,Kaid82}; in coherent gluon jet
emission (Cherenkov gluons, in particular)~\cite{Drem80}; in the cold
quark-gluon plasma model~\cite{VHov89}. Models of a still more
phenomenological nature have been tried, such as cluster
models~\cite{Fugl87,Fugl88,DrDu75,LiMe83}, clan models~\cite{GiVH88} and
narrow hadron jet emission~\cite{OchWo88-89}. Whereas in all these models
definite dynamical mechanisms are proposed for the origin of the fluctuations
in multiparticle production, they still suffer from one important
deficiency: they do not reveal the nature of the scaling laws observed for
factorial moments at small bin sizes.
{}From that point of view, one would prefer the random cascade models
{}~\cite{bialas1,bialas2,BiPe88} and/or the general approach of phase
transitions~\cite{AnCo90,Pesc89,CaSa87,DrNa91}. While the cascade
models rely on analogies with turbulence theories and lead to
phase transitions, general considerations of the transition from parton
to hadron phases of the process are based on important properties of strong
coupling field theories reminding of QCD lattice computations and the
conformal group symmetry. Both approaches lead in quite a natural way to
scaling behaviour of factorial moments and are preferred as heuristic tools.
However, to date they cannot compete with phenomenological Monte-Carlo
models in providing computational results comparable with experimental data
at the same level of precision. We shall discuss them separately at some
length later, together with important ideas of intermittency and fractality,
but first we shall describe a variety of phenomenological models applied
to the fluctuation problem.
First of all, we should mention the furthest developed Monte-Carlo versions
of QCD inspired models~\cite{LUND,DPM,QGSM,Mar88,Odo84,CaTr88,Ande87,Kaid82}
of parton showers or quark-gluon strings. Qualitatively, these models describe
the behaviour of the two-particle correlation function $C_2(y_1,y_2)$
observed experimentally, showing the signature of short-range correlations.
However, they cannot pretend to fit them quantitatively in hh collisions
for all topologies and cut-offs. All the models predict noticeably smaller
values of intermittency indices than the experimental data. Such models
also fail to describe the probability distribution of maximum particle
number per event at a given resolution $\delta y$~\cite{PeTr87,Ande88}.
As discussed in Subsect.~4.2.3, for $\re^+\re^-$-annihilation the situation is still
controversial. Initially, the DELPHI-collaboration, working at the $Z^0$
peak, claimed agreement with the LUND parton shower model. Later, increasing
statistics tenfold, it finds that the agreement only holds at the level
of 10-20\% (see Fig.~4.7d)~\cite{Abreu90}. However, in general the situation
is better here than in hadronic reactions. In nucleus-nucleus reactions,
these models fail to describe the damping of spectator particles observed
in experiment~\cite{ESt95}. Further, more detailed studies are needed.
Thus, we see that fluctuations appear to be a stumbling block for
phenomenological models. They meet with difficulties when confronted
with measured factorial moments, in particular in hadron-hadron collisions.
In earlier days, no data existed on factorial moments for small bins
and information on strong fluctuations in the number of particles
inside such bins existed only for individual events. A distinctive feature
of the fluctuations was the azimuthal symmetry of particles belonging to
the spike. The whole event showed a noticeable ring of particles in the
plane perpendicular to the collision axis. Precisely for this reason, the
very first attempt to explain such fluctuations was based on an analogy with
Cherenkov photon radiation.
The hypothesis of coherent emission of gluon jets~\cite{Drem80} (involving, in
particular, the Vavilov-Cherenkov mechanism) predicted that these jets should
be emitted in a narrow pseudo-rapidity bin at rather large angles in the
center-of-mass system of the colliding hadrons. All subsequent models did not
predict any particular polar angle dependence for the dense groups of produced
particles. This specific feature was experimentally verified in
pp-interactions at 205 and 360 GeV energies~\cite{DremSJNP90}. It
turned out that the distribution of centers of dense particle groups on the
pseudorapidity axis contained several peaks superimposed on a fairly strong
background. Consequently, the proposed mechanism of coherent jet
emission in hadron interactions probably does exist but is not dominant.
It could provide the only distinction~\cite{DremSJNP90} between pp and
$\Pp\Pap$ data available from ISR, but no analysis of that kind has yet been
performed. The ring-like events were observed in earlier cosmic-ray
experiments~\cite{Ale62}~-~\cite{Maru79} and in recent studies of
nucleus-nucleus collisions~\cite{OCher95}. Nevertheless, other mechanisms
must be involved since intermittency is also observed in $\re^+\re^-$-annihilation,
where the conditions for coherence seem unlikely to be satisfied.
Large fluctuations may arise in an extended blob of a cold quark-gluon plasma
{}~\cite{VHov89}. The appealing feature of such a model is the relationship
between this phenomenon and the production of soft and low-$p_\ifmath{{\mathrm{T}}}$ hadrons,
lepton pairs and photons. However, this model faces problems in explaining
the large values of intermittency indices in electron-positron annihilation
compared to hadronic and nuclear processes, since, contrary to present
experimental results, it leads one to expect that the effect is largest
in nucleus-nucleus collisions.
Models based on clusters~\cite{DrDu75,LiMe83} or clans~\cite{GiVH88} are
more flexible in fitting factorial moments, since they involve several
free parameters. It has been demonstrated~\cite{VHov89,Seib89} that the
existence of clans leads to a power-law increase of factorial moments for
smaller bin sizes. No quantitative comparison with experiment has been
attempted, however. As to cluster models, in some cases they succeeded
in describing the multiplicity distribution in symmetric rapidity bins
of various sizes~\cite{Fugl87,Fugl88} and of the two-particle correlation
function. The data available at that time were limited to rather large
rapidity interval sizes ($\delta y \geq 0.5$). In smaller regions, however,
simple cluster model fail completely (see Figs.~4.4b, 4.20c and 4.23a).
Multiparticle production is described somewhat differently
in~\cite{FoWe78,CaFrie89}, where particle emission is explained
by two types of sources - chaotic and coherent.
Some of the above approaches attempt to describe multiplicity distributions
in varying rapidity bins in terms of the negative binomial distribution
(or modifications thereof). In the cluster model, this is accomplished by
varying the cluster parameters. In the clan model, the negative binomial
distribution is obtained by compounding a Poisson distribution for the
number of clans with a logarithmic distribution for their decay. In the
statistical model with two types of sources, the negative binomial
distribution naturally describes the chaotic sources, while the coherent
sources contribute a Poisson component.
Even though the negative binomial distribution can be phenomenologically used
to fit experimental data in the very first approximation, there are definite
distinctions from it in experiment. Besides, the asymptotic QCD predictions
disfavour it revealing new features~\cite{Dr94} of multiplicity distributions.
In particular, the NBD cumulants are always positive, while perturbative
QCD predicts negative values (and oscillations) of the higher order
cumulants. The pQCD prediction is supported by experiment for the
case of the total rapidity range.
\subsection{Intermittency and fractality}
In most cases, the models considered above need to have their parameters
adjusted to be able to fit (if at all) the data on factorial moments.
Power-law behaviour of factorial moments is
most naturally obtained for cascading
mechanisms and in phase transitions, as we shall see later.
The concept of intermittency has been borrowed from the theory of turbulence.
There, it represents the following property of a turbulent fluid: vortices of
different size alternate in such a manner as to form a self-similar structure.
They do not fill in the whole volume, but form an intermittent pattern
alternating with regions of laminar flow.
Mathematically, this property is described by a
power-law dependence of the vortex distribution moments on the vortex size.
This is the reason why the exponents
$\phi_q$ in the power-law dependence of factorial
moments $F_q(\delta y)\propto (\delta y)^{-\phi_q}$ in (\ref{fq113}) are called
``intermittency indices''.
As mentioned in Sect.~2.4, the self-similar nature of vortices directly
implies a connection between intermittency and fractality. Fractals are
self-similar objects of a non-integral dimension. The fractal dimension is a
generalization of ordinary topological dimensionality to non-integers.
More complicated self-similar objects exist, consisting of differently
weighted fractals with different non-integer dimensions. They are called
multifractals and are characterized by generalized (or R\'enyi) dimensions
$D_q$ which depend on the rank $q$ of the moment of the probability
distribution over such objects. The analysis of multifractals according to
the L\'evy indices goes beyond the simple definition of intermittency.
The formal definitions are given in Sect.~2.4. For more details,
we refer to the review papers~\cite{Pala87,PescTH5891,DremSPU90} and
references therein. In connection with multiparticle production, fractals
were first mentioned in~\cite{Ven79,Gio79,Minh83,DremJETP87}.
Somehow, the concept of fractality goes beyond a purely formal definition of
intermittency, by connecting the observed dimensionality with the geometrical
and thermodynamical properties of an object, as well as with the properties of
the distributions over this object~\cite{Pala87,Hen83}. The power-law
behaviour of factorial moments reveals the fractal structure of rapidity
distributions as decomposed in individual events. Its relation to geometrical
and thermodynamical properties is discussed below.
In string models~\cite{LAnd}, the self-similar behaviour of fluctuations is
ascribed to the fractal nature of the phase space available for subsequent
branchings that is formulated in~\cite{Bj?} as a "plumber view" of
multiparticle processes.
Sometimes, intermittency is ascribed~\cite{ABi?} to the fluctuations of the
geometrical sizes of emitting sources.
Some caution is necessary when applying these concepts to multiparticle
production. As discussed in Chapter~4, finite statistics, the rather small
number of particles produced in an event and, especially, in a given cell,
the method of bin-splitting, the rather restricted range of bin widths over
which the power-law behaviour is observed, all influence the final
conclusion. This is described in more detail in~\cite{PescTH5891,LeTs91}.
\subsection{Random cascade models}
In turbulence, intermittency was first demonstrated in cascade
models~\cite{Kolm41}. Modifications of these, as applied to multiparticle
processes, are rather popular nowadays~\cite{bialas1,bialas2}.
In such models, one considers a series of self-similar steps in partitioning
phase-space. Let us denote by $M$ the number of bins obtained by breaking
up the total phase space into $\lambda$ parts at each of the $\nu$ iterations of
the self-similar cascade. Thus $M=\lambda^{\nu}$ ($\equiv\frac{\Delta Y}{\delta y}$
for a total rapidity range $\Delta Y$ divided into bins of width $\delta y$).
Random cascade models involve a probability distribution $r(W)$
with corresponding moments
\begin{equation}
\langle W^q\rangle = \int \ifmath{{\mathrm{d}}} W\ \ r(W)W^q\ ,\ \ \ \ \langle W\rangle = 1\ . \label{wq12}
\end{equation}
The function $r(W)$ induces density fluctuations as the rapidity window
is broken up into ever smaller bins. The density $P_m$ in the $m$-th bin
is given by the product
\begin{equation}
P_{m}=\frac{1}{M}\prod^\nu_{n=1} W_n\equiv\frac{1}{M}
\frac{\rho_{(m)}}{\langle \rho_{(m)}\rangle}, \label{pm13}
\end{equation}
where the sequence of indices $n$ defines a path leading to a given
bin $m$ with density $\rho_{(m)}$. One assumes that there exists a range
of scales inside of which the weights $W$ are constant, i.e., they do
not depend on the scale at which they operate.
Herefrom, the intermittent character of the models follows as:
\begin{equation}
F_q = \langle (MP_m)^q\rangle=\langle\prod^{\nu}_{n=1}W_{n}^{q}\rangle =
(\Delta Y/\delta y)^{\ln\langle W^q\rangle/\ln\lambda}\ . \label{fq14}
\end{equation}
The intermittency indices are equal to
\begin{equation}
\phi_q = \ln\langle W^q\rangle/\ln\lambda\ , \label{fq15}
\end{equation}
i.e., random cascade models possess a multifractal spectrum~\cite{Pala87}.
The simplest type of distribution $r(W)$ is the subclass of so-called
$\alpha$-models~\cite{Scher84} given by the two-level probability distribution:
\begin{equation}
r(W) = p\delta(W-W_-) + (1-p)\delta(W-W_+), \label{rw16}
\end{equation}
where $0\leq W_-<1<W_+$ and $pW_-+(1-p)W_+=1$ because of the normalization
condition (\ref{wq12}). The density enhancement $W_+>1$ occurs with
probability $(1-p)$ at each step of the cascade, while a depletion $W_-<1$ is
present with a probability $p$. Combined, they create ``spikes'' and
``holes'' in the rapidity distribution. The intermittency indices are given by
\begin{equation}
\phi_q = \ln[pW^q_-+(1-p)W^q_+]/\ln\lambda\ . \label{fq17}
\end{equation}
The study of moments and multifractal analysis reveal new interesting
features.
(Let us mention that the $\alpha$-model is reduced to the $\beta$-model for $W_-=0$
and describes a mono-fractal in that case). As the parameters $p$ and
$\lambda$ are changed, the model predicts various phase transitions
{}~\cite{PescTH5891,BiSZ90,BrPe90}. The moments of factorial moments are
useful to reveal these transitions due to the fact that the distributions of
factorial moments are extremely irregular by themselves~\cite{DremHolm90}.
Introducing the normalized moments of moments and ascribing to these
a power-law behaviour at small bins of the form
\begin{equation}
\langle Z^p_q\rangle = \frac{\langle F^p_q(\delta y)\rangle}{\langle F_q(\delta y)\rangle^{~^p}}\propto
(\delta y)^{p\phi_q} \langle F^p_q(\delta y)\rangle \propto (\delta y)^{\varepsilon_{p,q}} \ ,
\label{zp18}
\end{equation}
one can analyse, in the framework of $\alpha$-models, the dependence of the
indices $\varepsilon_{p,q}$ on the parameters of the model. As discovered
in~\cite{BiSZ90}, this dependence defines four regions in the parameter
space, which are reminiscent of four different phases. The indices
$\varepsilon_{p,q}$ act as order-parameters.
The same conclusions have been obtained when studying~\cite{BrPe90} the
normalized factorial correlators $F_{pq}/F_pF_q$. Another important property
of these correlators is their independence of the bin width
{}~\cite{bialas1,bialas2}. This property has been confirmed by experiments (see
Sect.~4.6).
However, the $\alpha$-model does not predict the correct dependence
on the distance between bins. It predicts power-law behaviour
with an exponent
\begin{equation}
\phi_{pq} = \phi_{p+q}-\phi_p-\phi_q \label{fp19}
\end{equation}
related to the usual intermittency indices. The experimental values
do not follow a straight line on a double-log plot. Moreover, there
are no finite intervals where they satisfy the above relation. When
roughly approximated by a straight-line fit, the experimental values
of $\phi_{pq}$ are larger than those of the $\alpha$-model.
However, one should keep in mind that this is a toy-model, which could
pretend to be valid for asymptotically long cascades, i.e. for
extremely high energies and multiplicities. Otherwise, one should
develop Monte-Carlo programs~\cite{LeTs91} losing the beauty of
analytical formulae. In fact, it has been clearly shown for such a
well-known mathematical model as the Cantor set~\cite{Levt90}, that it
is not an easy task to reveal its fractal dimension (known {\it a
priori}) using factorial moments, if the number of iterations is
finite.
Nevertheless, the heuristic value of $\alpha$-models in describing
qualitative features of the process is rather high since, in
particular, they suggest the possibility of phase transitions. The
nature of the transitions and their relation to the quark-hadron
transformation are not clear yet. An interesting observation is the
existence of ``non-thermal'' transitions similar to those in spin-glass
systems. They differ from the usual ``order-disorder'' transitions by
producing ``different order'' in different regions of phase space, so that
one may call them ``clustered order-disorder'' transitions. In that case,
the intermittency indices increase with increasing rank faster than
linear. Analogies to statistical-mechanics systems~\cite{Pesc89,BrPe90}
provide further insight into the nature of the transitions.
\subsection{Field-theoretical approach and phase transitions}
Besides the scenario of a self-similar cascade, as another extreme, a
higher-order quark-hadron phase transition has been proposed to
explain strong fluctuations leading to intermittency
patterns~\cite{AnCo90,DrNa91}. Evidently, the statistical mechanics
description is most useful here. Hadronization of a quark-gluon plasma
becomes the origin of exceptional events with large fluctuations
observed above a strong background of conventional events. Such a
point of view is supported by studies of a two-dimensional lattice of
Ising spins~\cite{Wosi89,Satz89}, which shows that intermittency
appears at the critical temperature. The intermittency indices are
directly related to critical exponents of the system.
Similar features have been found for the $q$-state Potts model on the
Bethe lattice ~\cite{Haj92} at the phase transition but without
long-distance correlation. Clear fractal structure is also observed
for $SU(2)$ gluodynamics near the phase-transition point in lattice
calculations ~\cite{Polik} .
The scenarios of cascading and of phase transitions need not
contradict each other, if one accepts the point of view that the role
of a quark-hadron transition is to fix the fractal pattern formed by
cascading. The fluctuations are ``frozen'' at the transition point and
can be computed by just considering this point.
A well-defined field-theoretical procedure exists to treat
fluctuations and phase transitions in common media~\cite{PatPok}. It
requires, first of all, the definition of an order-parameter and its
treatment as an effective fluctuation field. In case of multiparticle
production, one could choose the rapidity density distribution of
particles in individual events $\rho_{(\ifmath{{\mathrm{e}}})}(y)$ (or a function of it) as
an order-parameter which fluctuates about its inclusive average
$\rho(y)$ at each rapidity value $y$. Its function as an
order-parameter is clarified by the local parton-hadron duality
hypothesis, which has been successful in describing the experimental
data for electron-positron annihilation processes. This hypothesis
states that the average values of $\rho_{(\ifmath{{\mathrm{e}}})}(y)$ at partonic $\rho_{(\Pp)}$
and hadronic $\rho_{(\ifmath{{\mathrm{h}}})}$ levels differ by a (for all rapidities) common
numerical factor. In particular, if one defines~\cite{DrNa91} the
fluctuation field as
\begin{equation}
\varepsilon(y) = \frac{\rho_{(\ifmath{{\mathrm{e}}})}(y)}{\rho_{(\ifmath{{\mathrm{h}}})}(y)} -1 \ , \label{e20}
\end{equation}
then its average in the hadronic phase is equal to zero, while it
differs from zero at the partonic level:
$\langle\varepsilon(y)\rangle_{(\Pp)}=\rho_{(\Pp)}(y)/{\rho_{(\ifmath{{\mathrm{h}}})}(y)}-1=\,
\mbox{const}$.
Other possible choices for the fluctuation field exist (for example,
$\rho^{1/2}(y)$ ~\cite{AnCo90,CaSa87,ScSu73} or
$\rho-\rho_{(\ifmath{{\mathrm{h}}})}$~\cite{dnL92}). They have advantages and disadvantages
which we shall not discuss here.
The probability of a fluctuation is given by
\begin{equation}
W(\varepsilon)=Z^{-1}\exp[-F(\varepsilon)]\ , \label{we21}
\end{equation}
\begin{equation}
Z=\int \ifmath{{\mathrm{D}}}\varepsilon \ \exp[-F(\varepsilon)] \label{z22}
\end{equation}
where $\ifmath{{\mathrm{D}}}\varepsilon$ refers to the functional differential,
$F(\varepsilon)$ is the free energy and $Z$ is the partition function.
Adding to $F(\varepsilon)$ a term $J(y)\varepsilon(y)$ with an external
current $J(y)$, one obtains the irreducible Green functions:
\begin{equation}
\langle \varepsilon_1 \dots\varepsilon_q\rangle=\frac{\delta^q\ln Z}{\delta J(y_1)....\delta J(y_q)}
\vert_{J=0} \ , \label{ve23}
\end{equation}
which are related to correlation functions, so that, for example (we omit
the index h),
\begin{equation}
\langle\varepsilon_1\varepsilon_2\rangle=\frac{\rho_2(y_1,y_2)}{\rho(y_1)\rho(y_2)}-1
\equiv K_2(y_1,y_2)\ . \label{ve24}
\end{equation}
The factorial moments are easily obtained as
\begin{equation}
F_q(\delta y)=(\delta y)^{1-q}\int^{\delta y}_0 \ifmath{{\mathrm{d}}}\zeta_1...\int^{\delta y}_0 \ifmath{{\mathrm{d}}}\zeta_{q-1}
r_q(\zeta_1,\dots,\zeta_{q-1}) \label{fq25}
\end{equation}
where
\begin{equation}
\zeta_i=y_{i+1}-y_i \ \ and \ \ r_q=\rho_q(y_1,...,y_q)/\rho(y_1)...\rho(y_q)\ .
\label{zi26}
\end{equation}
At first sight, the fluctuation field theory is not directly related
to the underlying QCD.
Yet, these theories are connected through the fluctuation pattern of
individual events $\rho_{(\ifmath{{\mathrm{e}}})}(y)$, which should be described by both of
them if they pretend to be valid. Thus, our guesses on the fluctuation
field $\varepsilon(y)$ reflect special features of cascading and confinement in QCD.
For small fluctuations, one can represent $F(\varepsilon)$ by a Taylor series
\begin{equation}
F(\varepsilon) = F_0 + \int \ifmath{{\mathrm{d}}} y \left[ \frac{b}{2} \left[ \frac{\ifmath{{\mathrm{d}}}\varepsilon}{\ifmath{{\mathrm{d}}} y}
\right]^2
+ \frac{a}{2} \varepsilon^2 + c\varepsilon^3 + \ifmath{{\mathrm{d}}}\varepsilon^4 + \dots \right] \ , \label{fe27}
\end{equation}
which corresponds to the Ginzburg-Landau Hamiltonian when $c=0$,
$d\not=0$ and all higher terms are equal to zero. It has been
found~\cite{hwanaz} that some scaling indices (but not the critical
exponents) have universality properties in this approach.
For free fields, i.e. $c=d=\dots=0$, one gets
\begin{equation}
\langle\varepsilon_1\varepsilon_2\rangle_f = \gamma \exp \left[-|y_1-y_2|/\xi\right] \ , \label{e28}
\end{equation}
\begin{equation}
\gamma = \pi\xi/b; \ \ \ \xi=(b/a)^{1/2} \ . \label{g29}
\end{equation}
This exponential form fits the two-particle correlation function
qualitatively (and is often used, in particular, for
nucleus-nucleus collisions~\cite{Egg91}), but it does not provide
intermittency at small $\delta y$. One should remember that it is related
to the free-field Lagrangian but not to the Ginzburg-Landau potential
and, therefore, describes usual short-range correlations without any
phase transitions. One should also note that the approach is
formulated in momentum space and not in configuration space.
One is tempted to conclude that fluctuations are strong at small
rapidity intervals and that the perturbative approach fails. The
phenomenon of intermittency should be described by a strong coupling
field theory, where perturbative methods do not work. In particular,
the renormalization group approach and conformal theories have been
tried~\cite{DrNa91,dnL92} and have provided power-law behaviour of
Green functions and factorial moments at small bin widths. So, it
seems rather reasonable to fit the correlation function by an expression
\begin{equation}
C_2(y_1,y_2) \propto \frac{1}{|y_1-y_2|^\kappa} \exp [-|y_1-y_2|/\xi],\;\;\;\;
(\kappa<1)
\label{c229}
\end{equation}
for all rapidity separations. For rapidity separations smaller than
the correlation length $\xi$ one gets pure power-law dependence of the
correlation function due to a phase transition phenomenon. The
associated singularity should soften energy and transverse momentum
spectra. In the simplest approximation, intermittency indices increase
linearly with their rank.
Let us stress an important difference of the above consideration with
the previous treatment of phase transitions. The correlation length
$\xi$ does not tend to infinity and the exponential law is not
replaced by a power law at large $\delta y$, as one is accustomed
to. Instead, the power-law appears at small $\delta y$ ($\delta y \ll \xi $),
and does not influence the dependence at large rapidity. This
happens, because rapidities play now the role of coordinates in the
usual treatment, so that one has to deal with the ultraviolet (not
infrared) stable point of the Gell-Mann-Low function. One can
speculate that particles lying far apart on the rapidity scale reveal
the dynamics of the process with a finite correlation length related
to a particular form of the Lagrangian, while those at nearer points
remind of the self-organising critical processes with a scaling law
not tightly connected to a particular form of the Lagrangian (sandpile
phenomenon). Thus, correlations appear as ``{\it frozen} (due to
hadronization) {\it sounds}'' of cascading.
Similar problems have been discussed in the framework of
Feynman-Wilson fluid models~\cite{AnCo90,CaSa87}. One should introduce
the notion of temperature, additional assumptions on thermal
equilibrium, on Kadanoff scaling at the critical temperature, on the
relative role of conventional and stochastic (or quark-gluon plasma)
components, and so on. Imposing special boundary
conditions~\cite{acpv} on the grand-canonical partition function, one
can relate Kadanoff scaling in the fluid to KNO-scaling in
multiparticle processes and describe the fractal properties of the
fluid in a wide range of scales. Formula (\ref{c229}) appears to be
valid for correlation functions of the conventional hadronic system,
but for a system at the critical point, pure power-like behaviour with
a different exponent is restored. One is therefore lead to consider
the whole process in the framework of a two-component (conventional
$+$ critical) model.
The same approach has been extended~\cite{amd} to a multidimensional
analysis of intermittency using, however, the assumption that the
correlation functions factorize in rapidity and transverse momentum.
The predictions for factorial moments differ from QCD predictions.
The factorization hypothesis allows to proceed analytically and to
relate intermittency to fractal properties of the system in original
space-time (this problem has been addressed also
in~\cite{DremJETP87,dl92,Bial92}) but looks rather artificial for any
field theory (QCD included). For instance, if one assumes that
conformal symmetry is responsible for
intermittency~\cite{DrNa91,dnL92}, one obtains non-factorizable Green
functions and the predictions differ from the above-mentioned ones due
to the mixing of longitudinal and transverse momentum components,
inherent in field theories. We shall see, however, that conformal
theory and QCD also differ.
Numerical values of intermittency indices can be calculated in the
conformal scheme and agree qualitatively with experimental findings.
Again, the phase transition plays a crucial role.
Studies of the role of phase transitions in multiparticle production
are still in their infancy and have, until now, provided qualitative
results only. Also the relation between hadronization and phase
transitions in the simple cascade models treated in the previous
sub-section is not yet clear.
The relative role of parton cascading and hadronization is another
matter of debate. The problem would be solved if extreme proposals
were valid. Indeed, parton cascading with an ``infinite'' number of
steps would provide intermittency indices quadratically increasing
with their rank (this corresponds to ``wild'' or ``hard''
singularities). Phase transitions, on the other hand, yield
monofractal behaviour with a linear increase of the indices.
In reality, the two extremes must be modified
so that finite cascading
would lead to a slower increase, while the next operator-product
expansion terms for a phase transition
would induce a faster than linear rise of the intermittency exponent.
Such problems have as yet not been treated.
\subsection{The statistical-mechanics formalism}\\
Statistical analogy is a powerful way to analyse properties of
chaotic dynamical systems~\cite{Sinai72}, in general, and of
multifractals and cascade models,
in particular~\cite{Tel88,Pesc89,PescTH5891}.
It rests on the possibility to define a
partition function $Z(q)$ of the system in the following way:
\begin{equation}
Z(q) = \sum^M_{m=1} p_m^q \ , \label{zq30}
\end{equation}
where
\begin{equation}
p_m = \frac{\rho_{(m)}}{M\langle\rho_{(m)}\rangle} \label{pm31}
\end{equation}
is a normalized probability weight on the ensemble of bins $(m)$ and
$\rho_{(m)}$ is a random (rapidity) density registered in each bin $m$.
The relation to multifractal (or intermittent) properties of a system is
established if one considers systems for which the probability inside
a box $m$ is proportional to a power $\alpha_m$ of the box size and the number of
degenerate boxes (with the same value of $\alpha_m$) follows a power law as well.
Then, assuming a continuous limit, one finds as in (\ref{dr:121})
\begin{equation}
Z(q)\simeq \int\limits^{a_+}_{a_-} M^{f(\alpha)-q\alpha} \ifmath{{\mathrm{d}}}\alpha \ , \label{zq32}
\end{equation}
(integration running between maximum and minimum zeros of $f(\alpha)$),
wherefrom one easily obtains the multifractal spectrum of the system $f(\alpha)$
(see, for example,~\cite{Pala87,DremSPU90,DremFest}).
Remembering the definition of the intermittency indices $\phi_q$, one relates
them to the spectrum $f(\alpha)$ via the relation
\begin{equation}
f(\bar\alpha) = -q^2 \frac{\ifmath{{\mathrm{d}}}}{\ifmath{{\mathrm{d}}} q} \left[ \frac{\phi_q+1}{q}\right] \ ,
\label{fa}
\end{equation}
where $\bar\alpha$ is defined as
\begin{equation}
\bar\alpha(q) = 1 - \frac{\ifmath{{\mathrm{d}}}\phi_q}{\ifmath{{\mathrm{d}}} q}\ . \label{aq}
\end{equation}
The interpretation of $f(\alpha)$ is transparent since it weights the number of
degenerate boxes. It, therefore, corresponds to the entropy in statistical
mechanics. In a similar way, the rank $q$ may be interpreted as an inverse
``temperature'' $\beta=1/T$ and the relation (\ref{fa}) corresponds to the
usual thermodynamical formula
\begin{equation}
S = - \frac{dF}{dT}, \label{s35}
\end{equation}
where $S$ is the entropy and $F$ is the free energy, whose
``temperature'' dependence is provided now by
\begin{equation}
\lambda(q) = \frac{\phi_q+1}{q}=1-F\ . \label{laq}
\end{equation}
Two features of this analogy are particularly useful. On the one hand,
application of the thermodynamical formalism allows for coverage of
multifractals by boxes of different sizes, i.e. for a more precise
description of individual events. This has been used in proposals of
correlation measures with non-uniform
coverage~\cite{Hen83,Drem88,Alba90}. On the other hand, the
multifractal treatment admits an extension of the thermodynamical
formalism to non-equilibrium systems. It has been used to classify the
phase transitions in $\alpha$-models and to demonstrate that, in
multiparticle processes, phases could exist similar to spin-glass
states ~\cite{Pesc89,PescTH5891}.
It is important to note that the minimum of the $\lambda(q)$
(\ref{laq}) corresponds, according to (\ref{fa}), to zeros of the
fractal spectrum. This is a signal for a phase transition in
thermodynamical systems.
One should stress, however, that the similarity of the distributions
is in itself not sufficient to justify use of statistical physics
terminology in its original meaning for the quantum field systems we
are interested in here. Besides, the analogy breaks down completely
for values of $q$ exceeding the multiplicities effectively
contributing to the moments. Nevertheless, this analogy is used
in~\cite{Band,HeKr} to derive the dependence of the pressure in
a Feynman-Wilson liquid on its chemical potential and some peculiar
features are found in hadron-hadron reactions (see also the review
paper~\cite{DrLe95}).
It is evident that further explorations of statistical mechanics
approaches to multiparticle production are needed. In particular,
analytical properties of a partition function, often useful in
connection with phase transitions, have not been much analysed.
The location of the (complex) roots (zeros) $z\nu$ of the multiplicity
generating function (\ref{dr:15}) has recently been
studied in~\cite{zeros:edw,ID94} after earlier suggestions
by Biebl and Wolf~\cite{Brown}. This work is based on the
analogy with the famous Lee-Yang zeros\cite{Yan52},
whose location fully characterizes
the thermodynamic properties of the physical system.
For multiplicity distributions, the strength of the fluctuations of the
multiplicity in an event is directy related to
the location of the zeros in the complex $z$ plane:
the magnitude of the factorial cumulants, and thus the strength of the
correlations, is determined by the roots closest to the origin.
In the discrete version of QCD, developed in LUND, is was
demonstrated that the zeros of (\ref{dr:15}) belong to a fractal
Julia-set~\cite{zeros:bo} with intruiging properties.
Detailed studies of this set, and various connections with standard
phenomenology, such as KNO-scaling, remain to be worked out.
\subsection{Intermittency, evolution equations and QCD }
In the previous section we considered two rather extreme
possibilities, simple cascade models and phase transitions, as
possible mechanisms leading to scale-invariance in particle production
processes.
Further interesting results have been derived from studies of
simplified kinetic branching evolution equations for ``birth-death''
(or ``gain-loss'') processes. Many of these are treated in textbooks
{}~\cite{Bartl55,Sevas71} and were applied to multiparticle production
{}~\cite{Giovan79,LamWal84,Shih86,Batu88,Chliap90}. In general, the
time-evolution of the number of ``clusters'' (partons, resonances, etc.)
is described by forward or backward (retrospective) Kolmogorov
equations, which relate the time derivative of the generating function
to some combinations of that function. A particular example is the
Smoluchowski equation treated in ~\cite{Meun92}. The terms
``forward'' and ``backward'' imply, that each tree graph may be viewed
either as a splitting to ever ``thinner'' branches, or as a
convolution to ever ``thicker'' branches.
For linear evolution equations, the solutions depend directly on the
initial conditions~\cite{Biya90}. Non-linear equations often have
solutions asymptotically independent of
pre-history~\cite{Batu88,Meun92}. Stationary regimes may appear if
the annihilation of clusters is stronger than the ``birth-rate''. In
that case, the dispersion is proportional to the average multiplicity
and intermittent behaviour can be obtained, if the proportionality
factor is larger than one and the mean multiplicity decreases
correspondingly for smaller bins. This can easily be proven by means
of the definition of the second-order factorial moment.
Let us note that systems obeying non-linear evolution could exhibit
quite general properties, independent of the detailed form of the
equations, such as period-doubling. Some ideas along this line of
thought, using properties of stochastic systems and Feigenbaum
attractors, have been formulated in~\cite{Dias87,Batu92}.
More detailed analysis of intermittency in the framework of the
Smoluchowski equation~\cite{Meun92} reveals various regimes of
time-evolution and cascading, depending on the parameters of the
model. The Smoluchowski equation is of the backward type. It
contains terms linear and quadratic in the generating function $G$
with opposite signs. Formally it looks like
\begin{equation}
\frac{\ifmath{{\mathrm{d}}} G}{\ifmath{{\mathrm{d}}} t}=G*G - G*G_{1} \ , \label{dg37}
\end{equation}
where $G(u,t)=\sum_{n\geq 1} N(n,t)u^{n},\ \ \ G_{1}=G(1,t)$ with
$N(n,t)$ representing the number of clusters of (integer) mass
$n$ at time $t$ and the convolution $*$ is defined through the
aggregation coefficients of clusters. The fractal properties of
aggregates and the occurrence of phase transitions have been
analysed in~\cite{Meun92}.
Obviously, it would be desirable if an explanation of scaling phenomena
in multiparticle production could be derived from, so to say,
first principles, i.e. in the framework of QCD. Multiparticle production
in QCD is the result of quark-gluon branching and the subsequent transition
to hadrons. As such, the self-similar multiplicative branching (or cascade)
process could give rise to a scaling regime. The perturbative QCD parton
shower picture is justified for interactions with large transferred momenta
(or virtualities), but in hadronic reactions one mostly has to deal with
soft processes. Perturbative QCD is valid in the initial
stages of high-energy cascades in electron-positron annihilation
and could, therefore, be used as an explanation for intermittency.
It is well known that the perturbative QCD cascade gives rise~\cite{Alta77}
to a mean multiplicity of partons increasing rapidly with energy.
Equations for higher moments of parton multiplicity distributions are rather
complicated~\cite{Andre81,Dok92}, but reveal in any case that
the parton number distributions are much wider than a Poissonian. The infrared
limit becomes very important and one should consider infrared-safe properties.
Assuming that the singularity is avoided in a way similar
as in an electromagnetic cascade in a medium,
one can estimate the fractal dimension of internal
motion of partons in a jet and it turns out to be quite
low for a single jet in $\re^+\re^-$-annihilation~\cite{DremJETP87}.
The simplest theoretical models, such as the tree diagrams of
the $\phi^3$ model, simplified QCD~\cite{Hwa89,ChiuHwa89,Hwa90-91} also based
on tree diagrams, the Schwinger tunnelling transition~\cite{Bial89} or the
effective Lagrangian approach~\cite{ScSu73}, indicate that
the totality of all produced partons exhibits intermittency.
It is not yet clear what modifications of QCD equations going beyond
the tree graphs (so-called Double Logarithmic Approximation - DLA)
would fit this region best. An approach to that problem has been
proposed~\cite{GribLev} in deep inelastic processes where the transition
from the Bjorken limit to the Regge domain proceeds through some
intermediate region in which quadratic terms (in the fields) appear and
cause some recombination of partons at high densities. Here, the
evolution equation for the number of gluons $xG(x,q^{2})$
in a hadron with a transverse size $q^{-1}$ and for small values of
the Bjorken $x$-variable is taken to be
\begin{equation}
\frac{\partial^2 xG(x,q^2)}{\partial \ln (1/x) \partial \ln q^2}
=\alpha_sxG(x,q^2)-\frac{C\alpha_s^2}{q^2}
[xG(x,q^2)]^2 \ , \label{xg38}
\end{equation}
where $\alpha_{s}$ is the QCD coupling strength and $C$ is a
constant. It is inspired by QCD ideas and Regge phenomenology, but
has not been derived rigourously. No analysis of the intermittency
property has been attempted so far. One should note, however, that
the general effect of such a quadratic damping is to narrow the
multiplicity distribution (see, for example~\cite{Batu89}), which
leads to decreasing factorial moments.
One should, however, not rely on the similarity of
equations (\ref{dg37}) and (\ref{xg38}), since
this could be misleading. As is well known~\cite{Andre81,BrPe92,Dok92}, the
equations for the generating functionals for gluon and quark jets in QCD are
non-linear, while the corresponding equations for the structure
functions in DLA are just linear GLAP equations. The generating
functional for a parton p is given by
\begin{equation}
G_{\Pp}(u,v,x,Y)=\sum_{n_{\ifmath{{\mathrm{q}}}},n_{\ifmath{{\mathrm{g}}}}}\frac {1}{n_{\ifmath{{\mathrm{q}}}}!n_{\ifmath{{\mathrm{g}}}}!}\prod
\int \ifmath{{\mathrm{d}}} x_{\ifmath{{\mathrm{q}}},i} \ifmath{{\mathrm{d}}} x_{\ifmath{{\mathrm{g}}},j}u(x_i)v(x_j)W^{\Pp}_{n_{\ifmath{{\mathrm{q}}}}n_{\ifmath{{\mathrm{g}}}}}(x,x_{i},
x_{j},Y) \label{gp39}
\end{equation}
where $W$ is a differential probability to create $n_{\ifmath{{\mathrm{q}}}}$ quarks and
$n_{\ifmath{{\mathrm{g}}}}$
gluons with an evolution parameter $Y\sim \ln\ln Q^{2}$ in a p jet.
The equations for the generating functionals are
\begin{equation}
\frac{\partial G_{\ifmath{{\mathrm{q}}}}(x,Y)}{\partial Y}=\int_{0}^{1}\ifmath{{\mathrm{d}}} x' P_{\ifmath{{\mathrm{q}}}\rq}(x',x)
[G_{\ifmath{{\mathrm{q}}}}(x',Y)G_{\ifmath{{\mathrm{g}}}}(1-x',Y)-G_{\ifmath{{\mathrm{q}}}}(x,Y)], \label{gq40}
\end{equation}
\begin{eqnarray}
\frac{\partial G_{\ifmath{{\mathrm{g}}}}(x,Y)}{\partial Y}=& \displaystyle\int\limits_{0}^{1}\ifmath{{\mathrm{d}}} x'
[P_{\ifmath{{\mathrm{g}}}\rg}(x',x) G_{\ifmath{{\mathrm{g}}}}(x',Y)G_{\ifmath{{\mathrm{g}}}}(1-x',Y)-G_{\ifmath{{\mathrm{g}}}}(x,Y)) \nonumber\\
&+n_{f}P_{\ifmath{{\mathrm{q}}}\ifmath{{\mathrm{g}}}}(x',x)(G_{\ifmath{{\mathrm{q}}}}(x',Y) G_{\ifmath{{\mathrm{q}}}}(1-x',Y)-G_{\ifmath{{\mathrm{g}}}}(x,Y))] \ ,
\label{gg}
\end{eqnarray}
where the $P$'s are the corresponding GLAP kernels, $n_{\hrulefill}$ is the number of
flavours and the initial conditions are such that at $Y=0$, $G_{\ifmath{{\mathrm{q}}}}=u$ and
$G_{\ifmath{{\mathrm{g}}}}=v$ for a single jet.
The status of the equations for generating functionals is not
completely clear up to now. They are able to reproduce the higher
order graphs of the perturbation theory far beyond the tree
level~\cite{Dok92}. Their success in predicting the tiny features of
multiplicity distributions in total phase space (for a review
see~\cite{Dr94}) encourages speculations about their quite general
status, with some confinement properties taken into account already at
that stage.
A simplified version of (\ref{gg}) with the quark term omitted
(gluodynamics) has been studied in~\cite{BrPe92}. Intermittency for
the factorial moments of the gluon cascade is claimed to be in
qualitative agreement with experimental data, as well as evidence for
a structural phase transition. However, the formula for the
intermittency indices contradicts other QCD results. This is not
surprising, since the generating function technique should here
be applied to the subjet hitting the bin under investigation,
not to the whole jet as done in~\cite{BrPe92}.
The treatment of QCD cascades has been taken further in~\cite{Gust91}
within the framework of the dipole formalism including coherence effects.
The multifractal dimension of the parton cascade for high order q is
found to be equal to the QCD anomalous dimension
$\gamma_0=(6\alpha_{s}/\pi)^{1/2}$ and a first pre-asymptotic
correction has been calculated. Moreover, a geometrical interpretation
of the anomalous dimension of QCD is proposed.
A direct solution of QCD evolution equations has
been attempted for the second correlator in~\cite{Oc92}.
The behaviour of factorial moments of any rank (as well as of double
trace moments - see below) in small phase-space windows
for $\re^+\re^-$-collisions is treated both in DLA and in the next Modified Leading
Logarithmic Approximation (MLLA) of QCD in~\cite{dd92}.
Similar results for factorial moments in DLA are obtained
in~\cite{OW92a,BrMeuPe93}. They are closely related to the previously
derived formulae of~\cite{dmo}.
In the approach described above, one considers three stages of the process:\\
1. the initial quark emits a hard gluon,\\
2. the gluon evolves into a jet consisting of several subjets,\\
3. one of the subjets hits the phase space window chosen.\\
Integrating over all the stages one gets the final multiplicity distribution
(for details see~\cite{dd92}). For comparatively large windows one can use
the fixed coupling constant, while for smaller bins its running should be
taken into account.
Fixed coupling QCD factorial moments reveal~\cite{Oc92,dd92} the
intermittency phenomenon with intermittency indices equal to
\begin{equation}
\phi_{{\mathrm{QCD}}}(q)=(q-1)(D-\frac {q+1}{q}\gamma_{0}), \label{ddphi}
\end{equation}
in DLA. This formula is valid if, for the $D$-dimensional analysis with
$M$ bins along each axis, one defines $F_q \propto M^{\phi (q)}$.
For large $q$ the indices increase linearly. The term with negative
sign in the second bracket is proportional to the QCD multiplicity
anomalous dimension $\gamma_0$. From (\ref{ddphi}), one would conclude
that QCD prescribes fractal behaviour with codimension
\begin{equation}
\left.d_{q}\right|_{{\mathrm{QCD}}}=D-\frac {q+1}{q} \gamma_{0}. \label{ddcod}
\end{equation}
The phase space term $D$ is obviously non-fractal. The
$\gamma_{0}$-term in (\ref{ddcod}) is due to the energy dependence of
multiplicity and gives monofractal behaviour. The gluon energy
spectrum contribution, represented by $\gamma_{0} /q$, gives
multifractal behaviour. The next-to-leading corrections to
$\gamma_{0}$ also provide $q$-dependent terms.
The calculated values should be compared to the slopes in the region
$\delta y>1$ (which are rather large) since (\ref{ddcod}) is derived in
fixed-coupling QCD (the intermittent behaviour in that region was
discussed first in~\cite{SaSa}). In smaller bins, the QCD running
coupling becomes important and modifies the above relations
\cite{dd92}. The factorial moments now behave in a semi-power-like
manner so that there is no strict intermittency even though an
approximate one can still be claimed.
The second term becomes very close to 1 for the low-rank moments and
the low-rank intermittency indices turn out to be very small for
one-dimensional analyses, in accordance with experiment. Corrections
to $\gamma_{0}$ of the order of $\gamma_{0}^{2}$ can be taken into
account~\cite{dd92}. The influence of confinement seems negligible.
Thus small as well as large $\delta y$-intervals may be described in a
unified treatment, at least at a qualitative level.
The transition point from "large" to "small" bins depends on the rank
of the moment in QCD, in full accordance with experimental findings.
We should further mention that the ratio $d_{q}/d_{2}$ depends explicitly
on $D$, contrary to experimental claims (see Subsect.~4.3.3).
The ``free energy'' $F(q)$ is related to $\lambda (q)$ given by (\ref{laq}) as
\begin{equation}
F(q)=1-\lambda(q)\approx\gamma_0-\frac{\gamma_0}{q^2}\ \ .\label{lamb}
\end{equation}
In DLA, it is a steadily increasing function of $q$. However, with
corrections to $\gamma_{0}$ taken into account~\cite{dd92} $F(q)$ becomes
non-monotonic. This happens for rather large values of q, with the result,
that the distinction between factorial and usual moments becomes crucial and
statistical analogies inapplicable. These findings show the limitations of
perturbative QCD and provide further insight into the properties of
multiplicity distributions, such as KNO-scaling, in full phase
space~\cite{doka,dokb}.
In particular, QCD gives rise to the prediction~\cite{dokb,Dr94} of a
negative value of the cumulant of rank 5 confirmed by experiment and
to the general conclusion of a non-infinitely-divisible nature of total
multiplicity distributions in QCD (that prohibits, e.g., the one-ladder
multiperipheral cluster models).
The increasing branch of the multifractal spectrum $f(\alpha )$
may be easily calculated using (\ref{fa}), (\ref{aq}) giving
\begin{equation}
f(\alpha )=2\gamma_{0}^{1/2} (\alpha -\gamma_{0})^{1/2}. \label{ddfa}
\end{equation}
The double trace moments (DTM), redefined in analogy with factorial moments as
\begin{equation}
F_{q,\nu} \equiv \frac {1}{\Delta }\left(\sum_{\Theta_{m}\in \Delta}\frac
{n_{m}(n_{m}-1)...(n_{m}-\nu+1)}{n^{\nu}}\right)^{q}\ \ ,
\end{equation}
behave \cite{dd92} in QCD as
\begin{equation}
F_{\nu ,q} (\Delta ) \propto \Delta^{\frac {\gamma_{0} (q^{2}-1)}{q\nu }}
\propto \Delta ^{q\phi_{\nu }
-\phi_{q\nu } +q-1} \equiv \Delta^{-\phi (q,\nu )+q-1} , \label{DTMnuq2}
\end{equation}
wherefrom one finds
\begin{equation}
\phi (q,\nu )=\phi (q \nu )-q \phi (\nu ) = (q-1)\left(1-\frac {q+1}{q}\cdot
\frac {\gamma_{0}}{\nu }\right) . \label{pdif}
\end{equation}
The second factor in (\ref{pdif}) may be called ``double codimension''.
It is not symmetric in $q$ and $\nu$ and shows that increasing $\nu $
one decreases effectively the anomalous dimension. For $\nu =1$,
as required, the double codimension becomes equal to the usual codimension.
The scaling exponent (\ref{pdif}) is not factorizable in $\nu $ and $q$.
The above redefinition of DTM is aimed at reducing the Poissonian noise
and the role of phase-space factors, otherwise very important.
In fact, it is surprising that the above expression describes qualitatively
the general trends and even the absolute normalization of the
functions $K(q,\nu )$ shown in~\cite{Ratti91,Ratti92}.
For large $\nu$, the ratio $K(q,\nu )/(q-1)$
is completely determined by the phase space factor and
should tend to 1. This is seen in the experimental data.
In the region of small $q\sim 1$ and $\nu \sim 1$ the strong
compensation in (\ref{pdif}) prevents its use, but even there it
gives quite reasonable values of $K(q,\nu )$. This probably
indicates that DTM defined as powers of multiplicities are not
sensitive enough to dynamics, a suspicion raised earlier (Sect.~4.7).
On the other hand, the parton cascading picture employed
in~\cite{dd92} may be applied, strictly speaking, only to hard
processes at extremely high energies and one should not rely much on
the asymptotic estimates of QCD when considering experimental data.
Also, the difference between usual and factorial moments becomes
extremely important at large values of $q $ or at sufficiently small
bins with low multiplicities. One may suspect that relation
(\ref{pdif}) and, especially its first part, has a much wider range of
applicability than just for $e^{+}e^{-}$-collisions and some universal
relation could be valid for other reactions. If so, it will be
important to understand whether these common features are due to
common dynamics or to insufficient sensitivity of the proposed
measures.
The large fluctuations, {\it e.g.} those observed by the NA22 collaboration,
have raised the suspicion that at small bins one is dealing with unusually
wide distributions, which could have infinite moments. This has led one to
consider L\'evy-stable probability distributions. The L\'evy indices
derived from QCD factorial moment indices or DTM exponents show
\cite{dd92} no sign of ``wild'' singularities.
The DTM technique has been first applied to experimental data from
hadron-hadron reactions (described in Subsect.~4.7.7) where direct QCD
arguments are invalid. For $\re^+\re^-$-collisions, however, it may be
worthwhile to analyse the data with this method.
Let us further note that there is a difference between predictions
of QCD and those of variants of conformal theories considered
in~\cite{dnL92}, or of multiplicative models~\cite{amd}. For example,
the intermittency indices derived in a rapidity or azimuthal angle
analysis should be equal in QCD.
This is a consequence of the symmetrical form (in pseudorapidity $\eta
$ and azimuthal angle $\varphi $) of the gluon propagator, which can be
written as
\begin{equation}
k^{2}=(p_{\ifmath{{\mathrm{T}}},1}+p_{\ifmath{{\mathrm{T}}},2})^{2}=4p_{\ifmath{{\mathrm{T}}},1}p_{\ifmath{{\mathrm{T}}},2}(\sinh ^{2}
\frac{\eta_{12}}{2}+\sin^{2}\frac{\varphi_{12}}{2}) , \label{k2}
\end{equation}
where $p_{\ifmath{{\mathrm{T}}},i}$ is the transverse momentum of $i$-th parton.
In a conformal theory, the intermittency indices for the second factorial
moment are given by
\begin{equation}
\phi_2 (\delta y)=2\eta, \;\;\;\;\; \phi_2 (\delta \varphi )=0\ \ , \label{p2dn}
\end{equation}
respectively. Here, $\eta $ is the conformal anomalous dimension estimated
to be $\eta \approx 0.07-0.1$. For multiplicative models, one finds
\begin{equation}
\phi_2 (\delta y)=1-D_{y}, \;\;\;\; \phi_2 (\delta \varphi )=D_{\varphi }-1
\label{anp2}
\end{equation}
with $0 < D_{y} < 1, 1 < D_{\varphi } < 2$.
The emergence of intermittent behaviour in solutions of non-linear
equations and in perturbative QCD encourages further studies along
these directions. At the least, they hold a promise of further
theoretical insight. One should keep in mind, however, that a direct
comparison of QCD-based asymptotic results with present-day
experiments is not justified. Some effects revealed by the data seem
to be of a different, as yet unsatisfactorily explained, origin.
\chapter{Conclusions}
Developments in physics --- and in science in general ---
over the last decade, have brought exciting new discoveries and
deeper insight into the dynamics of complex systems.
Studies of classical and quantum chaos, non-equilibrium
dissipative processes, random media, growth phenomena and many
more have all contributed to reveal the pervasive importance
of self-similarity, of power-laws and of fractals in nature.
Research in these fields is still in full evolution and continues
to uncover intriguingly simple and often surprisingly
universal behaviour in complex, non-linear systems.
The suggestion of Bia\l as and Peschanski to look for
scaling in particle fluctuations was one of the first attempts
to apply modern ideas and techniques from complex-system
research to multihadron-production processes.
In preceding pages, we have presented a critical
overview of the impressive amount of experimental and theoretical
work this proposal has generated since its formulation in 1986.
The continuing interest in the field testifies of the
growing conviction that new avenues need to be explored
for progress in strong-interaction physics.
Impressive as it may be, this work has not yet led to
final answers concerning the fundamental issues.
Approximate power-law scaling of particle density and
correlation functions is now indeed observed,
especially in two- or three-dimensional phase-space.
However, so far it can be explained from an interplay between jet
formation and more or less ``conventional'' correlations
among identical particles due to quantum interference.
Nevertheless, as often happens, the detailed scrutiny of
data (and detectors) on the full variety of collision processes in the search
for power behaviour has led to many new observations
of interest in their own right.
It has helped to recognize the importance
of detailed studies of correlation phenomena at large and small
distances in momentum space and new sensitive and general techniques
have been developed for their analysis.
Standard hadronization models, all too often accepted as satisfactory,
have been exposed to severe and sometimes even painful tests.
Intermittency analysis has revealed deficiencies in our understanding
of the hadronization process. These defects are not easy to cure in a
consistent manner
by simple parameter-tuning and ``new'' physics may well be needed
to restore internal consistency in, e.g.,~fragmentation models of the LUND
type. Present work on this subject starts to provide hints
that purely probabilistic treatment of
the break-up of colour fields has to be supplemented
with effects deeply connected with the structure of the non-perturbative
QCD vacuum. Progress in this direction would in itself be ample
compensation for the efforts spent on attempting to establish
fractality in multihadron production.
Data obtained in the last years have shown the overwhelming importance of
correlations among identical particles in the ``intermittent'' regions
of phase space. This quantum mechanical phenomenon,
discovered in particle physics in 1959, still awaits
satisfactory incorporation into present hadronization phenomenology, if it
is to be used as a reliable interferometric tool, e.g.,~in
studies of quark-gluon plasma formation.
Theoretical work has developed along a large variety of directions.
Fractal properties have been discovered in string-fragmentation models.
Within the realm of perturbative QCD, parton correlations and
emergence of power behaviour are now studied with
increasing sophistication. The relevance for present-day
phenomenology remains doubtful, however.
The powerful methods of statistical mechanics have been intensively
exploited in studies of random cascades as well as in equilibrium and
non-equilibrium critical phenomena.
Results of real intrinsic value have been obtained, with potentially
interesting applications in other fields.
In search for an explanation of ``unusually large'' density
fluctuations, ``intermittency'', in analogy with fluid turbulence, has
progressively led to appreciate the importance and often spectacular
manifestation of non-linear strong-coupling dynamics.
Experiments in deep-inelastic scattering are now starting to probe
hadron structure in a regime where the perturbative parton-cascade
picture becomes blurred. Non-linear perturbative evolution and
confinement play an increasingly important role in the very low
Bjorken-$x$ region now accessible in HERA. Present attempts to
understand this region invoke QCD Reggeon-theory.
It is intriguing to speculate that a power-law dependence of the
low-$x$ parton correlation functions could manifest itself as ``gluonic''
intermittency in virtual parton cascades, with the occasional creation of
regions with very large, or very small gluon density.
\newpage
\subsubsection{Acknowledgements}\\
It is a great pleasure to thank the many physicists who have
contributed enthusiastically to this new field with experimental data,
theoretical or methodical ideas or constructive criticism. Among those
from which we have gained most considerably are
B.~Andersson,
I.V.~Andreev,
A.~de Angelis,
A.~Bia\l as,
J.~Bjorken, F.~Botterweck, B.~Buschbeck,
P.~Carruthers, A.~Capella, M.~Charlet,
S.~Chekanov,
H.~Dibon,
I.~Derado,
H.~Eggers,
E.L.~Feinberg,
K.~Fia\l kowski,
A.~Giovannini,
E.~Grinbaum-Sarkisyan,
G.~Gustafson,
R.~Hwa,
G.~Jancso,
V.~Kuvshinov,
A.V.~Leonidov,
P.~Lipa,
F.~Mandl,
H.~Markytan,
W.~Metzger,
J.-L.~Meunier,
W.~Ochs,
R.~Peschanski,
O.~Podobrin,
S.~Ratti,
I.~Sarcevic,
H.~Satz,
N.~Schmitz,
J.~Seixas,
E.~Stenlund,
A.~Syed,
F.~Verbeure,
R.~Weiner,
B.~Wo\v siek,
J.~Wo\v siek,
Y.F.~Wu,
K.~Zalewski, and many others to be found in the references.
It is a great favour to be able to work with these
scientists and to share their pioneering spirit in difficult territory.
|
\section{Introduction}
This paper continues a series of works on relativistic kinetic theory of
an $N$-body system \cite{HSP}--\cite{hadr}
within the framework of a manifestly covariant
mechanics \cite{HP}, both for the classical theory and the corresponding
relativistic quantum theory. In this framework, for the classical
case, the covariant dynamical evolution of a system of $N$ particles is
governed by equations of motion that are of the form of Hamilton
equations for the motion of $N\;events$ which generate the particle
space-time trajectories (world lines). These events are considered as the
fundamental dynamical objects of the theory and characterized by their
positions $q^{\mu }=(ct,{\bf q})$ and energy-momenta $p^{\mu }=(E/c,{\bf
p})$ in an $8N$-dimensional phase space. The motion is parametrized by a
continuous Poincar\'{e}-invariant parameter $\tau $ \cite{HP} called the
``historical time''. For the quantum case, the covariant dynamical
evolution of $N$ particles is governed by a generalized Schr\"{o}dinger
equation for the wave function $\psi _{\tau }(q_1,q_2,...,q_N)\in
L^2(R^{4N}),$ with measure $dq_1dq_2\cdots dq_N\equiv d^{4N}q,$
describing the distribution of events ${q_i\equiv q_i^{\mu },\;\mu =0,1,2
,3;\;i=1,2,\ldots ,N}.$ The collection of events (called
``concatenation'' \cite{AHL}) along each world line corresponds to a
$particle$ in the usual sense; e.g., the Maxwell conserved current is an
integral over the history of the charged event \cite{Jack}. Hence the
evolution of the state of the $N$-event system describes $a\;posteriori$
the history in space and time of an $N$-particle system.
The evolution of the system is assumed to be governed by Hamiltonian-type
equations with a Lorentz-invariant scalar function, the relativistic
dynamical function of the variables $(q_i,p_i)$ specifying the state of
each particle $i.$ In the simplest case of a free particle, for which the
world line is generated by a free event, the relativistic dynamical
function (generalized Hamiltonian) is $$K_0=\frac{p^\mu p_\mu }{2M},$$
where we use the metric $g^{\mu \nu }=(-,+,+,+),$ and $M$ is a given
fixed parameter (an intrinsic property of the event), with the dimension
of mass.
The Hamilton equations $$\frac{dq^\mu }{d\tau }=\frac{\partial K}
{\partial p_\mu },\;\;\;\frac{dp^\mu }{d\tau }=-\frac{\partial K}
{\partial q_\mu }$$ yield, in this case, $$\frac{dq^\mu }{d\tau }=\frac{p
^\mu }{M},\;\;\;\frac{dp^\mu }{d\tau }=0.$$ Eliminating $d\tau ,$ one
finds $$\frac{d{\bf q}}{dt}=\frac{{\bf p}}{E}c^2,$$ as required for the
motion of a free relativistic particle. It then follows that, for a free
motion, the proper time interval squared, divided by $d\tau ^2,$ is
$$\frac{dq^\mu }{d\tau }\frac{dq_\mu }{d\tau }=\frac{p^\mu p_\mu }
{M^2}.$$ For $$K_0=-\frac{M}{2},$$ corresponding to the ``mass-shell''
value $$p^\mu p_\mu =-M^2c^2,$$ it follows that $$c^2dt^2-d{\bf q}^2=c^2d
\tau ^2.$$ In the more general case in which $$K=K_0+V,$$ where $V$ is,
for example, a function of $q,\;\;p^2\equiv p^\mu p_\mu $ may vary from
point to point along the trajectory. Hence, in general, the proper time
interval does $not$ correspond to $d\tau .$
For a system of $N$ interacting events (and hence, particles) one takes
\cite{HP}
\beq
K=\sum _i\frac{p^\mu _ip_{i\mu }}{2M}+V(q_1,q_2,\ldots ,q_N),
\eeq
where all of the events are put, for simplicity, to have equal mass
parameters, and we write $q_i,$ for brevity, for the four-vector.
The Hamilton equations are
$$\frac{dq^\mu _i}{d\tau }=\frac{\partial K}{\partial p_{i\mu }}=\frac{p_
i^\mu }{M},$$
\beq
\frac{dp^\mu _i}{d\tau }=-\frac{\partial K}{\partial q_{i\mu }}=
-\frac{\partial V}{\partial q_{i\mu }}.
\eeq
These equations are precisely of the same form as those of
nonrelativistic Hamilton point mechanics, but in a space of $8N$
dimensions instead of $6N.$ The fundamental theorems of mechanics, such
as the Liouville theorem \cite{HSS}, the theory of
canonical transformations and Hamilton-Jacobi theory, follow in the same
way, with the manifold of space-time replacing that of space, and
energy-momentum replacing the momentum. It is fundamental to this
structure that there is a single universal evolution parameter $\tau $
which plays the role of the Galilean time.
In the quantum theory, the generalized Schr\"{o}dinger equation
\beq
i\hbar \frac{\partial }{\partial \tau }\psi _\tau (q_1,q_2,\ldots ,q_N)=K
\psi _\tau (q_1,q_2,\ldots ,q_N),
\eeq
with, for example, a $K$ of the form (1.1), describes the evolution of
the $N$-body wave function $\psi _\tau (q_1,q_2,\ldots ,q_N).$ To
illustrate the meaning of this wave function, consider the case of a
single free event. In this case, (1.3) has the formal solution $$\psi _
\tau (q)=(e^{-iK_0\tau /\hbar }\psi _0)(q)$$ for the evolution of the
free wave packet. Let us represent $\psi _\tau (q)$ by its Fourier
transform, in the energy-momentum space: $$\psi _\tau (q)=\frac{1}{(2\pi
\hbar )^2}\int d^4p\;e^{-ip^2\tau /2M\hbar }e^{ip\cdot q/\hbar }\psi _
0(p),$$ where $p^2\equiv p^\mu p_\mu,\;p\cdot q\equiv p^\mu q_\mu ,$ and
$\psi _0(p)$ corresponds to the initial state. Applying the Ehrenfest
arguments of stationary phase to obtain the principal contribution to
$\psi _\tau (q)$ for a wave packet centered at $p_c^\mu ,$ we find $$q_c^
\mu =\frac{p_c^\mu }{M}\tau ,$$ consistent with the classical equations
(1.2). Therefore, the central peak of the wave packet moves along the
classical trajectory of an event, i.e., the classical world line.
The wave functions have a local interpretation, i.e., $\vert \psi _\tau
(q)\vert ^2d^4q$ is the probability to find an event at the space-time
point $q^\mu $ in space-time volume $d^4q.$ Localization in space, as
well as in time, can be shown by applying arguments given in ref. \cite
{AH}.
Horwitz, Schieve and Piron \cite{HSP} have constructed equilibrium
classical and quantum Gibbs ensembles. They found that the grand
partition function in the rest frame of the system is given by
\beq
\ln Z(\beta ,V,\mu ,\mu _K)=e^{\beta \mu }\int \frac{d^4pd^4q}{(2\pi )^4}
e^{-\beta E}e^{\beta \mu _K\frac{m^2}{2M}},\;\;\beta =\frac{1}{k_BT}.
\eeq
In addition to the usual chemical potential $\mu $ in the grand canonical
ensemble, there is a new potential $\mu _K$ corresponding to the mass
degree of freedom of relativistic systems.
Horwitz, Shashoua and Schieve \cite{HSS} have shown that in the framework
of the manifestly covariant mechanics which we discuss here, covariant
Weyl transforms exist for observables, and therefore covariant
relativistic Wigner functions \cite{DK} can be constructed. In this way
they derived a manifestly covariant relativistic generalization of the
BBGKY hierarchy for the $s$-particle relativistic Wigner functions. By
approximating the effect of correlation of second and higher order by
two-body collision terms (using the cross-sections defined in ref.
\cite{HL}), as in the usual nonrelativistic Boltzmann theory, they
obtained a manifestly covariant Boltzmann equation (for non-identical
events). This equation was used to prove the $H$-theorem for evolution
in $\tau .$ In the equilibrium limit, a covariant form of the
Maxwell-Boltzmann distribution,
\beq
f^{(0)}(q,p)=e^{A(q)(p-p_c)^2},
\eeq
was obtained. Since this distribution
is the distribution of the $4$-momenta of the events, $m^2=-p^2=-p^{\mu }
p_{\mu }$ is a random variable in a relativistic ensemble. In order to
obtain a simple analytic result the authors restricted themselves to a
narrow mass shell $p^{2}=-m^{2}\cong -M^{2}.$ The results obtained in
this approximation are in agreement with the well-known results of Synge
\cite{Syn} for an on-shell relativistic kinetic theory.
In ref. \cite{di} the equilibrium Maxwell-Boltzmann distribution (1.5)
was considered for the whole range of $m,$ to obtain the corresponding
equilibrium relativistic $mass$ distribution. Its low-temperature and
nonrelativistic limits were investigated and shown to yield results in
agreement with nonrelativistic statistical mechanics \cite{galim}.
In the present paper we study the case of indistinguishable events.
In contrast to the approach of Horwitz, Shashoua and Schieve \cite{HSS},
we choose another approach initiated by Yang and Yao \cite{YY} in the
nonrelativistic case, which is based on the Wigner distribution
functions and the Bogoliubov hypotheses to find approximate dynamical
equation for the kinetic state of any nonequilibrium system \cite{Bog}.
Kinetic equation that we obtain, which represents a relativistic
generalization of the Boltzmann-Uehling-Uhlenbeck (BUU) equation
\cite{BUU} for indistinguishable particles, and can be easily generalized
to include the non-identical case as well. The generalized Boltzmann
equation obtained in this way is then used to prove the $H$-theorem for
evolution in $\tau .$ In the equilibrium limit, the covariant forms of
the Bose-Einstein/Fermi-Dirac/Maxwell-Boltzmann distributions are
obtained, which, as considered for the whole range of $m,$ provide the
corresponding equilibrium relativistic mass distributions. The
relativistic mass distributions are studied in the identical particle
case in \cite{ind}, and their possible consequences for high energy
physics and cosmology are considered, respectively, in \cite{hadr} and
\cite{therm}.
We introduce two-body interactions by taking the support of mutual
correlations for any two events to be in a relative $O(2,1)$-invariant
subregion of the full spacelike region, as done in the solution of the
two-body bound state problem \cite{AH1,AH2}, and for the extraction of
the partial wave expansion from the relativistic scattering amplitude
\cite{AH3}. We then calculate the expressions for the energy density and
pressure of an interacting gas, and show that they have the same form (in
terms of an invariant distance parameter) as those of the nonrelativistic
theory and provide the correct nonrelativistic limit.
\section{Relativistic $N$-body system}
The evolution in $\tau $ of an $N$-body system is determined by the
Liouville-von Neumann equation for the $N$-body density matrix $\rho $
(we use the system of units in which $\hbar =c=k_B=1,$ unless other units
are specified):
\beq
i\frac{\partial \rho }{\partial \tau }=[K,\rho ],
\eeq
where $K$ is the total $N$-body Hamiltonian, here taken to be
\beq
K=\sum _{i=1}^NK_i^{(0)}+\sum _{1=i<j}^NV_{i,j},
\eeq
where $$K_i^{(0)}=\frac{p_i^\mu p_{i\mu }}{2M}$$ and $$V_{i,j}=V\left( q_
i-q_j\right) ,\;\;\;q_i-q_j\equiv \sqrt {(q_i^\mu -q_j^\mu )(q_{i\mu }-q_
{j\mu })}$$ is a two-body interaction potential. In order to obtain the
BBGKY hierarchy, one introduces the $(n)$-body density matrices, as
follows:
\beq
\rho ^{(n)}_{1,2,\ldots ,n}=\frac{N!}{(N-n)!}Tr_{(n+1,\ldots ,N)}\rho ,
\eeq
\beq
Tr_{(1,2,\ldots ,n)}\rho ^{(n)}_{1,2,\ldots ,n}=\frac{N!}{(N-n)!},
\eeq
and, by taking the appropriate traces in Eq. (2.1), obtains \cite{BM}
\bqry
i\frac{\partial \rho ^{(n)}_{1,2,\ldots ,n}}{\partial \tau } & = &
\sum _{i=1}^n\;[K_i^{(0)},\rho ^{(n)}_{1,2,\ldots ,n}]\;+\sum _{1=i<j}^
n[V_{i,j},\rho ^{(n)}_{1,2,\ldots ,n}] \NL
& & +\;Tr_{(n+1)}\sum _{i=1}^n\;[V_{i,n+1},\rho ^{(n+1)}_{
1,2,\ldots ,n+1}].
\eqry
This set of equations is equivalent to (2.1).
In what follows, we shall use the simplified notation: $\rho _i\equiv
\rho _i^{(1)},\;\rho _{i,j}\equiv \rho _{i,j}^{(2)},$ etc., so that the
latter equation can be rewritten as
\beq
i\frac{\partial \rho _n}{\partial \tau }=\sum _{i=1}^n\;[K_i^{(0)},
\rho_n]+\!\sum _{1=i<j}^n[V_{i,j},\rho _n]+Tr_{(n+1)}\sum _{i=1}^n\;
[V_{i,n+1},\rho _{n+1}].
\eeq
It is convenient to introduce directly the symmetry requirements on the
function $\rho _n$ by means of
\beq
\rho _n=S_nF_n,
\eeq
where $S_n$ is a symmetrization/antisymmetrization operator defined by
\beq
S_n=\prod _{i=2}^n\left( 1\pm \sum _{j=1}^{i-1}P_{i,j}\right) .
\eeq
Here $P_{i,j}$ denotes the permutation operator. Since $S_n$ satisfies
the relation
\beq
S_{n+1}=S_n\left( 1\pm \sum _{i=1}^nP_{i,n+1}\right)
\eeq
and commutes with the operators $K_i$ and $V_{i,j},$ one can substitute
(2.7) into (2.6) and obtain the equation
\bqry
i\frac{\partial F_n}{\partial \tau } & = & \sum _{i=1}^n\;[K_i^{(0)},
F_n]\;+\sum _{1=i<j}^n[V_{i,j},F_n]\;+\;
Tr_{(n+1)}\sum _{i=1}^n\;[V_{i,n+1},F_{n+1}] \NL
& & \pm \;\;
Tr_{(n+1)}\sum _{i=1}^n\;[V_{i,n+1},\sum _{i=1}^nP_{i,n+1}F_{n+1}].
\eqry
Now we introduce the Wigner distribution functions \cite{DK},
\beq
f_s(q_s,p_s,\tau )=\frac{1}{(2\pi )^{4s}}\int dr_s\;F_s(q^{'}_s,q^{''}_s,
\tau )e^{ip_s\cdot r_s},
\eeq
\beq
F_s(q^{'}_s,q^{''}_s,\tau )=\int dp_s\;f_s(q_s,p_s,\tau )e^{
-ip_s\cdot r_s},
\eeq
where $$q^{'}_s=q_s-\frac12r_s,\;\;\;q^{''}_s=q_s+\frac12r_s,$$ and
$q_s\equiv (q_1,q_2,\ldots ,q_s),$ $p_s\equiv (p_1,p_2,\ldots ,p_s),$
$p_s\cdot r_s\equiv \sum _{i=1}^sp^\mu _ip_{i\mu },$ $dr_s\equiv dr_1dr_
2\cdots dr_s.$ One may substitute (2.12) into (2.10) and obtain the
quantum BBGKY hierarchy of the Wigner distribution functions $f_s=f_s(x_
s,\tau ),\;\;x_s=(q_s,p_s),$ as
\bqry
\frac{\partial f_s}{\partial \tau } & + & \sum _{j=1}^s\frac{p_j}{M}
\frac{\partial f_s}{\partial q_j}\;\;+\;\;i\sum _{j<k}^s\left( e^{i
\theta _{j,k}/2}-e^{-i\theta _{j,k}/2}\right) f_s \NL
& + & i\sum _{j=1}^s\int dx_{s+1}\;\left( e^{i\theta _{j,s+1}/2}-e^{-i
\theta _{j,s+1}/2}\right) f_{s+1} \NL
& \pm & i\sum _{j=1}^s\int dx_{s+1}\;\left( e^{i\theta _{j,s+1}/2}-e^{-i
\theta _{j,s+1}/2}\right) P_{j,s+1}f_{s+1}\;=\;0.
\eqry
Here $dx_s\equiv dq_sdp_s=d^4q_1\cdots d^4q_sd^4p_1\cdots d^4p_s,$ and
the operators
$\theta _{ij}$ and $\theta _{j,s+1}$ are represented as follows,
\beq
\theta _{ij}=\frac{\partial V_{ij}}{\partial q_i}\left( \frac{\partial }
{\partial p_i}-\frac{\partial }{\partial p_j}\right) ,\;\;\;\theta _{j,s+
1}=\frac{\partial V_{ij}}{\partial q_i}\frac{\partial }{\partial p_j}.
\eeq
For $s=1$ and 2, one finds
\bqry
\frac{\partial f_1}{\partial \tau } & + & \frac{p_1}{M}\frac{\partial
f_1}{\partial q_1}\;\;+\;\;i\int dx_2\;\left( e^{i\theta _{1,2}/2}-e^{-i
\theta _{1,2}/2}\right) f_2 \NL
& \pm & i\int dx_2\;\left( e^{i\theta _{1,2}/2}-e^{-i
\theta _{1,2}/2}\right) P_{1,2}f_2\;=\;0, \\
\frac{\partial f_2}{\partial \tau } & + & \sum _{j=1}^2\frac{p_j}{M}
\frac{\partial f_2}{\partial q_j}\;+\;i\left( e^{i\theta _{
1,2}/2}-e^{-i\theta _{1,2}/2}\right) f_2 \NL
& + & i\sum _{j=1}^2\int dx_3\;\left( e^{i\theta _{j,3}/2}-e^{-i
\theta _{j,3}/2}\right) f_3 \NL
& \pm & i\sum _{j=1}^2\int dx_3\;\left( e^{i\theta _{j,3}/2}-e^{-i
\theta _{j,3}/2}\right) P_{j,3}f_3\;=\;0.
\eqry
Equations (2.15) and (2.16) are exact. Since $f_2$ depends on $f_3,$
accurate solution of the hierarchy is impossible. One has, therefore, to
apply some approximated approach. One of such approaches is the
Bogoliubov one \cite{Bog}, which we shall apply in the present
consideration.
According to the Bogoliubov hypotheses \cite{Bog},
1) It is possibe to find a kinetic state of any non-equilibrium system,
provided that the average interval between two subsequent collisions is
much longer than the duration of the collision. In this kinetic state,
\beq
f_s(x_1,\ldots ,x_s;\tau )=f_s(x_1,\ldots ,x_s|f_1),
\eeq
\beq
\frac{\partial f_1}{\partial \tau }=A(x|f_1).
\eeq
2) There are no correlations in the initial state of a system. One can
introduce the displacement operator,
\beq
{\cal P}^s_\tau f_s(x^0_1,\ldots ,x_s^0)=f_s(x_1,\ldots ,x_s),
\eeq
where $x_1^0,\ldots ,x_s^0$ are the values of each $x$ at $\tau =0,$ and
$x_1,\ldots ,x_s$ are their values at $\tau .$ The non-correlative
condition at the initial state implies
\beq
{\cal P}^s_{-\tau }\left[ f_s(x_1,\ldots ,x_s)-\prod _{1\leq j\leq s}
f_1(x_j)\right] \rightarrow 0.
\eeq
Starting from the Bogoliubov hypotheses, it is possible to derive a
kinetic equation.
Although the invariant interaction potential has infinite support in
space-time, since it depends on $({\bf q}_1-{\bf q}_2)^2-c^2(t_1-t_2)^2,$
its long-range part is necessary close to the light cone. It has been
shown \cite{HS}, that wave operators exist in scattering theory if the
support of the wave function does not extend to zero mass. The space-time
volume $v$ of the effective interaction is therefore bounded. We shall
assume here that it may be taken to be small, as in the first hypothesis
of Bogoliubov. One can, therefore, write
\beq
\frac{\partial f_1}{\partial \tau }=A^0(x|f_1)+vA^1(x|f_1)=\ldots ,
\eeq
\beq
f_s=f_s^0+vf_s^1+v^2f_s^2+\ldots .
\eeq
In the first-order approximation, one sets
\beq
f_2\cong f_2^0\cong f_1(1)f_1(2)
\eeq
(henceforth we use the notation $1\equiv (x_1;\tau ),\;2\equiv (x_2;\tau
),$ etc.) and finds from (2.15)
\bqry
\frac{\partial f_1(1)}{\partial \tau } & + & \frac{p_1}{M}\frac{
\partial f_1(1)}{\partial q_1}\;\;+\;\;i\int dx_2\;\left( e^{i\theta
_{1,2}/2}-e^{-i\theta _{1,2}/2}\right) f_1(1)f_1(2) \NL
& \pm & i\int dx_2\;\left( e^{i\theta _{1,2}/2}-e^{-i
\theta _{1,2}/2}\right) P_{1,2}f_1(1)f_1(2)\;\;=\;\;0.
\eqry
This self-consistent equation is a relativistic generalization of the
quantum Vlasov equation \cite{YY}.
In the second-order approximation, one writes a formal solution,
\beq
f_s(x_1,\ldots ,x_s|f_1)=\sum _{i<j\leq s}g(x_i,x_j)\prod _{
k\neq i\neq j}f_1(k),
\eeq
where
\beq
g(x_i,x_j)=f_2^1(x_i,x_j|f_1)
\eeq
is a two-body correlation function, whose boundary condition is
\beq
\lim _{\tau \rightarrow \infty }{\cal P}_{-\tau }^{(2)}g(x_i,x_j)=0.
\eeq
Eq. (2.25) means that $s$-body effects are correlated by two-body
effects. One can write
\bqry
\frac{\partial f_2}{\partial \tau }\;=\;\frac{\partial f_2}{\partial
f_1}\frac{\partial f_1}{\partial \tau } & \approx & \left( \frac{
\partial f_2^0}{\partial f_1}+v\frac{\partial f_2^1}{\partial f_1}\right)
\left[ A^0(x|f_1)+vA^1(x|f_1)\right] \NL
& \approx & D_0f_2^0\;+\;v\left[ D_0\;g(x_1,x_2)+D_1f_2^0\right] ,
\eqry
where $$D_0\equiv A^0\frac{\partial }{\partial f_1},\;\;\;
D_1\equiv A^1\frac{\partial }{\partial f_1}.$$ One now uses Eqs.
(2.16),(2.25) and obtains
\bqry
D_0\;g(x_1,x_2) & + & \sum _{j=1}^2\frac{p_j}{M}\frac{\partial }{\partial
q_j}g(x_1,x_2)\;\;+\;\;i\sum_{j=1}^2\left( e^{i\eta _j/2}-e^{-i\eta _j/2}
\right) g(x_1,x_2) \NL
= & - & i\left( e^{i\theta ^{'}_{1,2}/2}-e^{-i\theta ^{'}_{1,2}/2}\right)
f_1(1)f_1(2)\;\;-\;\;i\int dx_3\left( e^{i\theta ^{'}_{1,3}/2}-e^{-i
\theta ^{'}_{1,3}/2}\right) \NL
& & \times \;g(x_2,x_3)f_1(1)\;\;-\;\;i\int dx_3\left( e^{i\theta
^{'}_{2,3}/2}-e^{-i\theta ^{'}_{2,3}/2}\right) f_1(2)g(x_1,x_3) \NL
& \mp & i\int dx_3\left( e^{i\theta ^{'}_{1,3}/2}-e^{-i\theta ^{'}_{1,
3}/2}\right) f_1(1)f_1(2)f_1(3) \NL
& \mp & i\int dx_3\left( e^{i\theta ^{'}_{2,3}/2}-e^{-i\theta ^{'}_{2,
3}/2}\right) f_1(1)f_1(2)f_1(3).
\eqry
Once $g(x_1,x_2)$ is known, one can obtain
the two-order-approximated equation for $f_1:$
\bqry
\frac{\partial f_1(1)}{\partial \tau } & + & \frac{p_1}{M}\frac{
\partial f_1(1)}{\partial q_1}\;\;+\;\;i\left( e^{i\eta _1/2}-e^{-i\eta
_1/2}\right) f_1(1)\;\;+\;\;i\int dx_2\;\left( e^{i\theta
_{1,2}^{'}/2}-e^{-i\theta _{1,2}^{'}/2}\right) \NL
& & \times \;g(x_1,x_2)\;\;\pm \;\;i\int dx_2\;\left( e^{i\theta
_{1,2}^{'}/2}-e^{-i\theta _{1,2}^{'}/2}\right) f_1(1)f_1(2)\;\;=\;\;0.
\eqry
Here $$\theta ^{'}_{1,2}=\frac{1}{v}\theta _{1,2},\;\;\theta ^{'}_{1,3}=
\frac{1}{v}\theta _{1,3},\;\;\eta _1=\frac{\partial U_1}{\partial q_1}
\frac{\partial }{\partial p_1},$$ and
\beq
U_1(q_1,\tau )=\frac{1}{v}\int dx_2\;f_1(2)V(q_1-q_2)
\eeq
is the mean-field potential.
In general, it is very difficult to obtain simultaneously solutions of
Eqs. (2.29) and (2.30). In the following section we show how Eq. (2.29)
can be solved for a quasihomogeneous system.
\subsection{Quasihomogeneous system}
The condition on a quasihomogeneous system is
\beq
g(x_1,x_2)=g(q_1-q_2,p_1,p_2)\equiv g(q,p_1,p_2),
\eeq
i.e., the correlation function depends only on the relative coordinates.
In this case, one obtains a formal solution for $g(q,p_1,p_2)$ by means
of the displacement techniques, as follows:
\bqry
g(q,p_1,p_2) & = & \int _0^\infty d\tau \;\left[ i\left\{ \left( e^{
\frac{i}{2}\frac{\partial }{\partial q}(\frac{\partial }{\partial p_1}-
\frac{\partial }{\partial p_2})}-e^{-\frac{i}{2}\frac{\partial }{\partial
q}(\frac{\partial }{\partial p_1}-\frac{\partial }{\partial p_2})}\right)
V\left( q-\frac{p_1-p_2}{M}\tau \right) \right\} \right. \NL
& & \times f_1(1)f_1(2)\;+\;i\int dq^{'}dp_3\;\left(
e^{\frac{i}{2}\frac{\partial }{\partial q}\frac{\partial }{\partial
p_1}}-e^{-\frac{i}{2}\frac{\partial }{\partial q}\frac{\partial }{
\partial p_1}}\right) V\left( q-q^{'} \right. \NL
& & \left. -\;\frac{p_1-p_2}{M}\tau \right) \times \left(
g(q^{'},p_2,p_3)f_1(1)\;\pm \;f_1(1)f_1(2)f_1(3)\right) \NL
& & \pm \;i\int dq^{'}dp_3\;\left(
e^{\frac{i}{2}\frac{\partial }{\partial q}\frac{\partial }{\partial
p_2}}-e^{-\frac{i}{2}\frac{\partial }{\partial q}\frac{\partial }{
\partial p_2}}\right) V\left( q-q^{'}-\frac{p_1-p_2}{M}\tau \right) \NL
& & \left. \times \left( g(q^{'},p_1,p_3)f_1(2)
\;\pm \;f_1(1)f_1(2)f_1(3)\right) \right] \NL
& = & i\int d\tau \;\left[ \left( e^{i\theta ^{'}_{1,2}/2}-e^{-i\theta
^{'}_{1,2}/2}\right) f_1(1)f_1(2)\right. \NL
& & +\;\int dx_3\;\left( e^{i\theta ^{'}_{1,3}/2}-e^{-i\theta ^{'}_{1,
3}/2}\right) \times \Big( g(x_2,x_3)f_1(1)\;\pm
\;f_1(1)f_1(2)f_1(3)\Big) \NL
& & \left. \pm \;\int dx_3\;\left( e^{i\theta ^{'}_{2,3}/2}-e^{-i\theta
^{'}_{2,3}/2}\right) \times \Big( g(x_1,x_3)f_1(2)\;\pm
\;f_1(1)f_1(2)f_1(3)\Big) \right] . \NL
& &
\eqry
In order to solve Eqs. (2.30) and (2.33), it is convenient to introduce
the Fourier transform, as follows:
\beq
\tilde{g}(k,p_1,p_2)=\int dq\;g(q,p_1,p_2)e^{-ik\cdot q},
\eeq
\beq
\tilde{V}(k)=\int dq\;V(q)e^{-ik\cdot q}.
\eeq
Substituting Eqs. (2.34),(2.35) into (2.30), one finds
\beq
\frac{\partial f_1}{\partial \tau }+\frac{p_1}{M}\frac{\partial f_1}{
\partial q_1}+F\frac{\partial f_1}{\partial p_1}=-\frac{i}{(2\pi )^4}\int
dk\;\left( e^{\frac{k}{2}\frac{\partial }{\partial p_1}}-e^{-\frac{k}{2}
\frac{\partial }{\partial p_1}}\right) \tilde{V}_{1,2}(k)h(k,p_1),
\eeq
where
\beq
h(k,p_1)=\int dp_2\;g(k,p_1,p_2),
\eeq
\beq
F\frac{\partial f_1}{\partial p_1}=i\left( e^{i\eta _1/2}-e^{-i\eta _1/2}
\right) f_1(1)\pm i\int dx_2\;\left( e^{i\theta ^{'}_{1,2}}-e^{-i\theta
^{'}_{1,2}}\right) f_1(1)f_1(2).
\eeq
Making a Fourier transform of Eq. (2.33), one obtains, after some
manipulations,
\bqry
{\rm Im}\;h(k,p_1) & = & \int dp_2\;\frac{\pi \tilde{V}_{1,2}(k)}{k|1\mp
\tilde{V}_{2,3}\psi |^2}\left[f_1^+(1)f_1^-(2)-f_1^+(2)f^-_1(1)\right]\NL
& & \times \;\delta \left( k\cdot \frac{p_1-p_2}{M}\right) .
\eqry
Here
\beq
f^{\pm}=f\left( p\pm \frac{k}{2}\right) \left[ 1\pm f\left( p\mp
\frac{k}{2}\right) \right]
\eeq
(the second sign $\pm $ in (2.40) distinguishes between bosons and
fermions),
\beq
f\left( p\pm \frac{k}{2}\right) =e^{\pm \frac{k}{2}\frac{\partial }{
\partial p}}f(p),
\eeq
and
\beq
\psi =\int _{-\infty }^\infty \frac{dp_3}{k\cdot (\frac{p_1-p_2}{M})-i
\varepsilon }\left[ f_1^+(3)-f_1^-(3)\right] .
\eeq
Substituting (2.39) into (2.36), one finally obtains
\bqry
\frac{\partial f_1(1)}{\partial \tau } & + & \frac{p_1}{M}\frac{
\partial f_1(1)}{\partial q_1}\;\;+\;\;F\frac{\partial f_1(1)}{
\partial p_1}\;\;=\;\;\frac{\pi }{(2\pi )^4}\int dk\;\left( e^{\frac{k}{
2}\frac{\partial }{\partial p_1}}-e^{-\frac{k}{2}
\frac{\partial }{\partial p_1}}\right) \NL
& & \times \int dp_2\;\delta \left( k\cdot \frac{p_1-p_2}{M}\right)
\frac{\tilde{V}_{1,2}^2(k)}{|1\mp \tilde{V}_{2,3}\psi |^2}\left[ f_1^+(
1)f_1^-(2)-f_1^+(2)f_2^-(1)\right] . \NL
& &
\eqry
Equation (2.43) is the kinetic equation of a gas of indistinguishable
particles in the quasihomogeneous case (the improved BUU equation
\cite{YY}). It reduces to the usual BUU equation provided that the
many-body effects are neglected and that the first-order approximation
for the term $F$ is taken:
\bqry
\frac{\partial f_1(1)}{\partial \tau } & + & \frac{p_1}{M}\frac{
\partial f_1(1)}{\partial q_1}\;\;-\;\;\frac{\partial U_1}{\partial
q_1}\frac{\partial f_1(1)}{\partial p_1}\;\;=\;\;\frac{\pi }{(2\pi )^{
12}}\int dp_2dp_1^{'}dp_2^{'}\;\delta ^{4}(p_1+p_2-p_1^{'}-p_2^{'}) \NL
& & \times \;\Big| \langle p_1p_2|V_{1,2}|p_1^{'}p_2^{'}\rangle \Big|
^2\left\{ f_1(1^{'})f_1(2^{'})\left[ 1\pm f_1(1)\right] \left[ 1\pm
f_1(2)\right] \right. \NL
& & \left. -\;\;f_1(1)f_1(2)\left[ 1\pm f_1(1^{'})\right] \left[
1\pm f_1(2^{'})\right] \right\} .
\eqry
In contrast to the usual Boltzmann and BUU equations which are applicable
in the restriction on the system to be dilute, Eq. (2.44) includes the
influence of many-body effects. Therefore, Eq. (2.44) provides an
essential improvement for the systems that have a higher particle density
or a larger force range of particle interaction; e.g., for strongly
interacting matter, heavy-ion collisions, or a cold relativistic plasma.
Rewriting this equation in the form
\bqry
\frac{\partial f_1(1)}{\partial \tau } & + & \frac{p_1}{M}\frac{
\partial f_1(1)}{\partial q_1}\;\;-\;\;\frac{\partial U_1}{\partial
q_1}\frac{\partial f_1(1)}{\partial p_1}\;\;=\;\;\frac{\pi }{(2\pi )^{
12}}\int dp_2dp_1^{'}dp_2^{'}\;\delta ^{4}(p_1+p_2-p_1^{'}-p_2^{'}) \NL
& & \times \;\Big| \langle p_1p_2|V_{1,2}|p_1^{'}p_2^{'}\rangle \Big|
^2\left\{ f_1(1^{'})f_1(2^{'})\left[ 1+\sigma f_1(1)\right] \left[ 1+
\sigma f_1(2)\right] \right. \NL
& & \left. -\;\;f_1(1)f_1(2)\left[ 1+\sigma f_1(1^{'})\right] \left[
1+\sigma f_1(2^{'})\right] \right\} ,\;\;\;\sigma =\pm 1,
\eqry
one sees that it reduces to the usual Boltzmann equation for
non-identicai particles for $\sigma =0.$ Thus, the three cases,
$$\sigma =\left\{ \begin{array}{rl}
1, & {\rm Bose-Einstein}, \\
-1, & {\rm Fermi-Dirac}, \\
0, & {\rm Maxwell-Boltzmann},
\end{array} \right. $$ can be treated by means of a unique equation,
(2.45), which can be, therefore, called the generalized Boltzmann
equation.
\section{Boltzmann $H$-theorem}
We now wish to establish the relativistic analogue to the Boltzmann
$H$-theorem and to prove that the entropy of an ensemble of events,
evolving without external disturbances, is nondecreasing as a function of
$\tau .$
The density of states in phase space associated with the distribution $n$
has been found in \cite{HSP},
$$\triangle \Gamma (\bar{n})=\left\{ \begin{array}{ll}
(\bar{n}+g-1)!/\bar{n}!(g-1)!, & {\rm Bose-Einstein} \\
g!/\bar{n}!(g-\bar{n})!, & {\rm Fermi-Dirac} \\
g^{\bar{n}}/\bar{n}! & {\rm Maxwell-Boltzmann}
\end{array} \right. $$ where $g$ is a number of states in each elementary
cell of energy-momentum space (degeneracy) and $\bar{n}$ is the average
occupation number. Assuming no degeneracy $(g\rightarrow 1)$ and using
Stirling's approximation $$\ln N!\approx N\ln N,\;\;\;N>>1,$$ we obtain
for the density of entropy in phase space, $s,$ $$\frac{s}{k_B}\equiv
\ln \triangle \Gamma (\bar{n})=\left\{ \begin{array}{ll}
-\bar{n}\ln \bar{n}+(1+\bar{n})\ln \;(1+\bar{n}), & {\rm Bose-Einstein}
\\ -\bar{n}\ln \bar{n}-(1-\bar{n})\ln \;(1-\bar{n}), & {\rm Fermi-Dirac}
\\ -\bar{n}\ln \bar{n}, & {\rm Maxwell-Boltzmann}
\end{array} \right. $$
\beq
=(\sigma +\bar{n})\ln \;(1+\sigma \bar{n})-\bar{n}\ln \bar{n},\;\;\;
\sigma =\pm 1,0.
\eeq
Therefore, in the case we are considering, the entropy of the ensemble
is defined by the functional
\beq
\frac{S(\tau )}{k_B}=\int dqdp\;\Big[ (\sigma +f_1(q,p;\tau ))\ln \;(1+
\sigma f_1(q,p;\tau ))-f_1(q,p;\tau )\ln f_1(q,p;\tau )\Big] .
\eeq
Then, taking the derivative of $S(\tau )/k_B,$ using Eq. (2.45) and
integration by parts of the space-time derivatives, we obtain, after some
manipulations,
\bqry
\frac{1}{k_B}\frac{dS}{d\tau } & = & \frac{\pi }{4(2\pi )^{12}}\int dqdp_
1dp_2dp_1^{'}dp_2^{'}\delta ^4(p_1+p_2-p_1^{'}-p_2^{'})\Big| \langle
p_1p_2|V_{1,2}|p_1^{'}p_2^{'}\rangle \Big| ^2 \NL
& & \times f_1(q,p_1;\tau )f_1(q,p_2;\tau )f_1(q,p_1^{'};\tau )f_1(q,p_
2^{'};\tau)\left[ \left( \frac{1}{f_1(q,p_1;\tau )}+\sigma \right)
\right. \NL
& & \left. \times \left( \frac{1}{f_1(q,p_2;\tau )}+\sigma \right)
\;-\;\left( \frac{1}{f_1(q,p_1^{'};\tau )}+\sigma \right)
\left( \frac{1}{f_1(q,p_2^{'};\tau )}+\sigma \right) \right] \NL
& & \times \left[
\ln \left\{ \left( \frac{1}{f_1(q,p_1;\tau )}+\sigma \right)
\left( \frac{1}{f_1(q,p_2;\tau )}+\sigma \right) \right\} \right. \NL
& & -\;\left. \ln \left\{ \left( \frac{1}{f_1(q,p_1^{'};\tau )}+\sigma
\right) \left( \frac{1}{f_1(q,p_2^{'};\tau )}+\sigma \right) \right\}
\right] .
\eqry
In the derivation of (3.3) the principle of microscopic irreversibility
(e.g., detailed balance) $$\Big| \langle p_1p_2|V_{1,2}|p_1^{'}
p_2^{'}\rangle \Big| ^2dp_1dp_2=\Big| \langle p_1^{'}p_2^{'}|V_{1,2}|p_1
p_2\rangle \Big| ^2dp_1^{'}dp_2^{'}$$ and the hermiticity condition
$$\Big| \langle p_1p_2|V_{1,2}|p_1^{'}p_2^{'}\rangle \Big| ^2=\Big|
\langle p_1^{'}p_2^{'}|V_{1,2}|p_1p_2\rangle \Big| ^2$$ were used.
Since $\Big| \langle p_1p_2|V_{1,2}|p_1^{'}p_2^{'}\rangle \Big| ^2
\delta ^4(p_1+p_2-p_1^{'}-p_2^{'})\geq 0,$ and the remaining factor in
the integrand is non-negative, we obtain
\beq
\frac{dS(\tau )}{d\tau }\geq 0,
\eeq
the relativistic Boltzmann $H$-theorem.
This result implies that the entropy $S(\tau )$ is monotonically
increasing as a function of $\tau ,$ and hence the evolution of the
system, as described by the generalized relativistic Boltzmann equation,
is irreversible in $\tau ,$ but {\it not necessarily in t.} In a smooth
average sense, one can argue that the entropy must increase in $t$ as
well. The support of the distribution function in $t$ is finite at each
$\tau ;$ as $\tau $ increases, this supprort moves up the $t$-axis, since
the system as a whole moves with the free motion of the center of mass.
The entropy, according to the $H$-theorem in $\tau ,$ must therefore also
be nondecreasing, in this coarse-grained sense, in $t.$ In the
nonrelativistic limit \cite{HR} $t\rightarrow \tau ,$ $S(t)$ takes on
the usual nonrelativistic form, and the nonrelativistic $H$-theorem for
evolution in $t$ is recovered.
In the special case in which the ensemble consists of positive energy (or
negative energy) states alone, a precise $H$-theorem can be proved for
the Lyapunov function $$\frac{\tilde{S}(t)}{k_B}=\int \!d^3q\;d\tau \!
\int _{p^0>0}\!\!\!d^4p\;\frac{p^0}{M}\Big[ (\sigma +f_1(q,p;\tau ))\ln
\;(1+\sigma f_1(q,p;\tau ))-f_1(q,p;\tau )\ln f_1(q,p;\tau )\Big] ,$$ by
the application of the arguments contained in ref. \cite{HSS}.
\subsection{Relativistic four-momentum distributions}
As we have seen in the preceding section, the entropy (3.2) of a system
of events increases, according to the generalized relativistic Boltzmann
equation, monotonically in $\tau .$ It means that the momentum
distribution function monotonically approaches its equilibrium value $f_1
^{(0)}(q,p).$ The equilibrium limit is achieved when
\beq
\frac{dS(\tau )}{d\tau }=0.
\eeq
Since the integrand in (3.3) is definite, (3.5) requires that, for the
equilibrium distribution $f_1^{(0)}(q,p),$
\beq
\ln \left( \!\frac{1}{f_1^{(0)}(q,p_1)}+\sigma \!\right)
+\;\ln \left( \!\frac{1}{f_1^{(0)}(q,p_2)}+\sigma \!\right) =
\ln \left( \!\frac{1}{f_1^{(0)}(q,p_1^{'})}+\sigma \!\right)
+\;\ln \left( \!\frac{1}{f_1^{(0)}(q,p_2^{'})}+\sigma \!\right) ;
\eeq
this condition implies the vanishing of the collision term in the
generalized relativistic Boltzmann equation (2.45).
Since $p_1,p_2$ and $p_1^\prime ,\;p_2^\prime $ are the initial and
final four-momenta for any scattering process, the general solution of
(3.6) is of the form
\beq
\ln \left( \frac{1}{f_1^{(0)}(q,p)}+\sigma \right) =\chi _1(q,p)+
\chi _2(q,p)+\ldots ,
\eeq
where the $\chi _i$ exhaust all quantities for which
\beq
\chi _i(q,p_1)+\chi _i(q,p_2)
\eeq
are conserved in collisions. The quantities conserved in the sense of
(3.8) are the individual event four-momentum $p^\mu $ and mass squared $m
^2\equiv -p^2$ (the latter is asymptotically conserved in the scattering
process \cite{HL}), and a constant (the one-event ``angular
momentum'' $M^{\mu \nu }=q^\mu p^\nu -q^\nu p^\mu $ also satisfies this
requirement, but does not change the structure of the result). Hence, the
most general form of $f_1^{(0)}$ is given by \cite{HSS,di}
\beq
\ln \left( \frac{1}{f_1^{(0)}(q,p)}+\sigma \right) =-A(p-p_c)^2-B,
\;\;\;\;A=A(q),\;B=B(q),
\eeq
where $p^\mu p_{c\mu }$ is an arbitrary linear combination of the
components $p^\mu ,$ so that
\beq
f_1^{(0)}(q,p)=\frac{1}{e^{-A(p-p_c)^2-B}-\sigma }=\left\{
\begin{array}{ll}
\frac{1}{\exp \{-A(p-p_c)^2-B\}-1}, & {\rm Bose-Einstein,} \\ & \\
\frac{1}{\exp \{-A(p-p_c)^2-B\}+1}, & {\rm Fermi-Dirac,} \\ & \\
e^{A(p-p_c)^2+B}, & {\rm Maxwell-Boltzmann.}
\end{array} \right.
\eeq
The physical properties of the distributions (3.10) are studied in
\cite{di} for the case of non-identical particles, and in \cite{ind} for
the case of identical particles. We shall normalize these distributions
as (the physical meaning of such a normalization is manifested below):
\beq
\int dqdp\;f_1^{(0)}(q,p)=V^{(4)},
\eeq
where $V^{(4)}$ is the total four-volume occupied by the ensemble in
space-time. Let us introduce the system of the space-time densities, as
follows:
\bqry
\int dp\;f_1^{(0)}(q,p) & \equiv & n_1^{(0)}(q), \\
\int dp_1dp_2\;f_2^{(0)}(q_1,q_2,p_1,p_2) & \equiv &
n_2^{(0)}(q_1,q_2), \\
\int dp_1dp_2dp_3\;f_3^{(0)}(q_1,q_2,q_3,p_1,p_2,p_3) & \equiv &
n_3^{(0)}(q_1,q_2,q_3), \;\;\;\;{\rm etc.}
\eqry
Then the one-body density, $n_1^{(0)}(q),$ is normalized, in view of
(3.11), as
\beq
\int dq\;n_1^{(0)}(q)=V^{(4)}.
\eeq
In the case of no $q$-dependence of $A$ and $B,$ Eq. (3.13) yields $n_1
^{(0)}=1.$
\section{Mean-field potential. RMS}
In the equilibrium case, Eq. (2.31) for the mean-field potential entering
the generalized Boltzmann equation (2.45), reduces to
\beq
U_1(q_1)=\frac{1}{v}\int dq_2dp_2\;f_1^{(0)}(q_2,p_2)V(q_1-q_2).
\eeq
Averaging (4.1) over the ensemble gives, through (3.11)--(3.13),
\bqry
U & \equiv & \frac{1}{2V^{(4)}}\int dq_1dp_1\;f_1^{(0)}(q_1,p_1)U_1(
q_1) \NL
& = & \frac{1}{2V^{(4)}v}\int dq_1dp_1dq_2dp_2\;f_1^{(0)}(q_1,p_
1)f_1^{(0)}(q_2,p_2)V(q_1-q_2) \NL
& \cong & \frac{1}{2V^{(4)}v}\int dq_1dq_2\;n_2^{(0)}(q_1,q_2)V(q_1-q_
2),
\eqry
where we have used the relation $f_2\approx f_1(1)f_1(2).$
The total energy density of the ensemble is defined by
\beq
\rho =\rho _0+\rho _{int},
\eeq
where $\rho _0$ is the energy density of a free gas (no-interaction case)
calculated in refs. \cite{di,ind}, and $\rho _{int}$ is the contribution
of the interaction potential which is equal to $$\rho _{int}=N_0U,$$
$N_0$ being the particle number density per unit comoving ``proper''
three-volume $V^{(3)},$ $N_0=N/V^{(3)}.$ We now assume that for the
interacting gas $V^{(4)}/N\sim v;$ it then follows form (4.2) that
\beq
\rho =\rho _0+\frac{N^2}{2(V^{(4)})^2V^{(3)}}\int dq_1dq_2\;n_2^{(0)}
(q_1,q_2)V(q_1-q_2).
\eeq
For a quasihomogeneous system, $n_2^{(0)}(q_1,q_2)=n_2^{(0)}(q_1-q_2),$
so that (4.4) takes on the form
\beq
\rho =\rho _0+\frac{N^2}{2V^{(4)}V^{(3)}}\int dq\;n_2^{(0)}(q)V(q).
\eeq
By introducing hyperbolic variables for spacelike $q,$
$$q^0=q\sinh \beta ,\;\;\;q^1=q\cosh \beta \sin \theta \cos \phi ,$$
\beq
q^2=q\cosh \beta \sin \theta \sin \phi ,\;\;\;q^3=q\cosh \beta \cos
\theta ,
\eeq
$$0\leq q<\infty ,\;\;-\infty <\beta <\infty ,\;\;0\leq \theta \leq \pi ,
\;\;0\leq \phi <2\pi ,$$ one can rewrite the integral in Eq. (4.5) as
$$4\pi \int q^3dq\cosh ^2\beta d\beta \;n_2^{(0)}(q)V(q).$$ This integral
does not have, however, a simply interpretable nonrelativistic limit, as
we discuss below after Eq. (4.12).
Let us instead turn to ref. \cite{AH1}, where the two-body
relativistic quantum-mechanical bound-state problem has been studied. It
was found that, if the support of the wave function of the relative
motion is restricted to an $O(2,1)$-invariant subregion of the full
spacelike region, one finds a lower mass eigenvalue of the ground state
than in the case when the support is in the full spacelike region. The
solutions, moreover, have a simply interpretable nonrelativistic limit.
This subregion was called by the authors the ``restricted Minkowski
space'' (RMS). It has a parametrization (in contrast to (4.6)
corresponding to the full spacelike region) $$q^0=q\sin \theta \sinh
\beta ,\;\;\;q^1=q\sin \theta \cosh \beta \cos \phi ,$$
\beq
q^2=q\sin \theta \cosh \beta \sin \phi ,\;\;\;q^3=q\cos \theta .
\eeq
Clearly, $q_1^2+q_2^2-q_0^2=q^2\sin ^2\theta \geq 0$ (and $q_1^2+q_2^2+
q_3^2-q_0^2=q^2\geq 0$ as well). This submeasure space is
$O(2,1)$-invariant, but not $O(3,1)$-invariant. The representations of
$O(3,1)$ are induced from the irreducible representations of $O(2,1)$
which are provided by the eigenfunctions of the two-body bound-state
problem \cite{AH2}. The fact that this restricted subregion admits a
lower mass of the ground state than the full spacelike region constitutes
a spontaneous symmetry breaking of the $O(3,1)$ invariance of the
dynamical equations.
The restriction of the relative coordinates to the RMS corresponds to a
restricted range of correlations available to the two events
propagating in a bound state, i.e., to the range of $q^\mu _1-q^\mu _2$
available at each $\tau .$ In computing the full spectrum of the two-body
problem the authors assumed that the wave functions of the excited states
also lie in the $O(2,1)$-invariant subregion, i.e., these correlations
are maintained for excited states as well. Indeed, it was found that the
partial wave expansion for scattering theory is recovered in this
submeasure space as well \cite{AH3}. Here we shall assume that this
result has more generality and can be applied in statistical mechanics:
{\it for any two events, their mutual correlations lie in the relative}
$O(2,1)$-{\it invariant subregion of the full spacelike region.}
It then follows that the two-body density $n_2^{(0)}(q_1-q_2)$ will have
support lying in the RMS associated with the relative motion $q^\mu _1-q^
\mu _2.$ Therefore, the integral in Eq. (4.5) will be nonzero only in
the RMS associated with $q,$ according to the nonvanishing support of the
two-body density $n_2^{(0)}(q).$ In this way we obtain for the integral
in Eq. (4.5)
\beq
\int q^3dq\sin ^2\theta d\theta \cosh \beta d\beta d\phi \;n_2^{(0)}(q)
V(q).
\eeq
We shall also assume that, at any instant of $\tau ,$ the extent of the
ensemble in the $q^0$-direction is bounded \cite{HSP}, so that $V^{(4)}=
V^{(3)}\cdot \triangle t,$ where $\triangle t$ is the range of the time
variable for the system as a whole. Therefore, in Eq. (4.8) $-\frac{
\triangle t}{2}\leq q^0\leq \frac{\triangle t}{2},$ and integration on
$\beta $ gives $$\int _{-{\rm Arc}\!\sinh (\triangle t/2q\sin \theta )}^
{{\rm Arc}\!\sinh (\triangle t/2q\sin \theta )}\cosh \beta d\beta =
\frac{\triangle t}{q\sin \theta };$$ Eq. (4.8) then reduces to
\beq
\triangle t\int q^2dq\sin \theta d\theta d\phi \;n_2^{(0)}(q)V(q)=
4\pi \triangle t\int dq\;q^2n_2^{(0)}(q)V(q).
\eeq
Using now the relations $V^{(4)}=V^{(3)}\cdot \triangle t$ and $N/V^{(3)}
=N_0,$ the particle number density, one finally obtains from (4.5),(4.9)
\beq
\rho =\rho _0+\frac{N_0^2}{2}\int d^3q\;n_2^{(0)}(q)V(q),
\eeq
where $d^3q$ stands for $4\pi q^2dq.$
In the same way it is possible to obtain the expression for the pressure
of the interacting gas \cite{therm}:
\beq
p=p_0-\frac{N_0^2}{6}\int d^3q\;q\frac{dV(q)}{dq}n_2^{(0)}(q).
\eeq
We see that the expressions for $\rho $ and $p$ are precisely of the same
form as those of nonrelativistic statistical mechanics \cite{Fey}, but
with $q\equiv \sqrt{q^\mu q_\mu }$ replacing $r\equiv \sqrt{{\bf q}^2},$
and $V(q)$ replacing $V(r).$ The situation is quite similar to the one
occuring in the two-body bound-state problem \cite{AH1}, where, upon
separation of variables in the RMS, one is left with a radial equation
for $q\equiv \sqrt{q^\mu q_\mu }$ which is of the same form as a
nonrelativistic radial Schr\"{o}donger equation for $r\equiv \sqrt{{\bf q
}^2}.$ Separation of variables in the RMS therefore has a clear
correspondence to the nonrelativistic problem, as first remarked in
\cite{AH1}. In the nonrelativistic limit, the relative variables $q^0$
and $p^0$ vanish (all the particles are synchronized in this limit
\cite{HSS}), and the formulas (4.10),(4.11) acquire their standard
nonrelativistic expressions \cite{Fey}.
We remark that the integral in Eq. (4.5) in the full spacelike region
can be made convergent in the same way, by imposing the bounds on the
time variable, as follows: $-\triangle t/2\leq q\sinh \beta \leq
\triangle t/2.$ In this case integration on $\beta $ results in the
expression $$4\pi \left( \frac{\triangle t}{2}\right) ^2\int dq\;
qV(q)n_2^{(0)}(q),$$ and we obtain, in place of (4.10),
\beq
\rho =\rho _0+\frac{N_0^2}{2}\;\frac{\triangle t}{4}\;
4\pi \int dq\;q\;n_2^{(0)}(q)V(q),
\eeq
and similar relation for $p.$ Hence, apart from $T_{\triangle V},$ the
average passage interval in $\tau $ for the events which pass through a
small representative four-volume of the system \cite{HSS}, contained in
the expressions for $\rho _0,p_0$ and $N_0$ \cite{di,ind}, there will
be another $T_{\triangle V}$ entering the expressions for $\rho _{int}$
and $p_{int},$ upon replacement for $\triangle t$ in the corresponding
formulas, through the relation (which represents the averaging of the
equation of motion for $q^0,$ $dq^0/d\tau =p^0/M,$ over the ensemble,
$\langle E\rangle $ being the average energy) $$\triangle t=T_{\triangle
V}\frac{\langle E\rangle }{M}.$$ In the nonrelativistic (or in the
sharp-mass) limit, $T_{\triangle V}\rightarrow \infty ,$ which provides a
stationarity of the system in space-time, but not a non-trivial evolution
in $\tau $ \cite{jpa}. While $p_0,\rho _0$ and $N_0$ are preserved in
this singular limit, due to the relation \cite{jpa} $T_{\triangle V}
\triangle m=2\pi ,$ where $\triangle m$ is the width of the mass
deviation from its on-shell value, $p_{int}$ and $\rho _{int}$ turn out
to converge with $T_{\triangle V}.$ Therefore, Eq. (4.12) and similar
formula for $p$ do not have a well-defined nonrelativistic limit, in
contrast to (4.10),(4.11), which admit its clear form. This fact should
be a source of a spontaneous symmetry breaking of the $O(3,1)$ invariance
in the correlation function of a many-body problem.
We remark that no problem with the convergence of the integral in Eq.
(4.5) arises in 1+2 dimensions (for the extent in the $q^0$-direction
bounded). Indeed, the 3D analog of (4.5) reads
\beq
\rho =\rho _0+\frac{N^2}{2V^{(3)}V^{(2)}}\int d^3q\;n_2^{(0)}(q)V(q).
\eeq
By introducing hyperbolic variables for spacelike $q,$
\beq
q^0=q\sinh \beta ,\;\;\;q^1=q\cosh \beta \sin \theta ,\;\;\;
q^2=q\cosh \beta \cos \theta ,
\eeq
$$0\leq q<\infty ,\;\;-\infty <\beta <\infty ,\;\;0\leq \theta \leq
2\pi ,$$ we rewrite the latter integral as $$2\pi \int q^2dq\cosh \beta d
\beta \;n_2^{(0)}(q)V(q).$$ Integration on $\beta $ gives $$\int _{-{\rm
Arc}\!\sinh (\triangle t/2q)}^{{\rm Arc}\!\sinh (\triangle t/2q)}\cosh
\beta d\beta =\frac{\triangle t}{q};$$ therefore, one obtains, via
$\triangle t=V^{(3)}/V^{(2)},$ $N/V^{(2)}=N_0,$
\beq
\rho =\rho _0+\frac{N_0^2}{2}\int d^2q\;n_2^{(0)}(q)V(q),
\eeq
the 3D analog of (4.10) ($d^2q$ stands for $2\pi qdq),$ and a similar
relation for $p.$
\section{Concluding remarks}
We have generalized the nonrelativistic approach of Yang and Yao, based
on the Wigner distribution functions and the Bogoliubov hypotheses, to
the relativistic case. We have derived the generalized Boltzmann equation
which, in the case of indistinguishable particles, improves the standard
BUU equation in three main aspects:
1) The effect of Pauli blocking, $f\rightarrow f(1-f^{'}),$ is included
in the collision term. This is important for the collision processes at
intermediate and low temperature, e.g., in heavy-ion collisions.
2) The modified mean-field interaction is introduced into the collision
term. This has a great influence on far-nonequilibrium states.
3) The equation takes into account binary collisions corrected for
many-body effects, wherein the many-body shielding effect can be obtained
spontaneously.
We have introduced two-body interactions, by means of the direct action
potential $V(q),$ where $q$ is an invariant distance in the Minkowski
space. The two-body correlations are taken to have the support in a
relative $O(2,1)$-invariant subregion of the full spacelike region, in
order to provide a good nonrelativistic limit to the basic thermodynamic
quantities. Since the expressions for the energy density and the pressure
are identical in form to those of the nonrelativistic theory, some of the
results for the nonrelativistic interacting gas should be applicable for
an interacting off-shell gas as well. For example, the equation of state
of the ideal gas of non-identical particles is \cite{di} $p=N_0T;$
therefore, it follows from (4.10),(4.11) that the equation of state of a
relativistic interacting gas should have the same form (by methods
analogous to those of the standard cluster expansion \cite{Hua}) as that
of a similarly interacting nonrelativistic one, i.e. \cite{Hua},
$$\frac{p}{N_0T}=\sum _{l=1}^\infty a_l(T)\left( \lambda ^3N_0\right) ^{
l-1},$$ where $\lambda \equiv \sqrt{2\pi /MT}$ is the thermal
wavelength, and $a_l(T)$ is the $l$th virial coefficient ($a_1=1).$
Applications of the generalized Boltzmann equation to realistic physical
systems, e.g., heavy-ion collisions, are now under consideration.
\section*{Acknowledgments}
We wish to thank R.I. Arshansky for discussions concerning the notion of
the restricted Minkowski space (RMS) and its role in relativistic
physics.
\bigskip
\bigskip
|
\section{Introduction}
\renewcommand{\theequation}{1.\arabic{equation}}
\setcounter{equation}{0}
Among the various conformal field theories, the Wess-Zumino-Witten
(WZW) models \cite{Wit,KZ,GW} take a somewhat special position. First
of all, due to the vast knowledge about the representation theory of
the affine Kac-Moody algebras \cite{PS,Kac,GO}, their mathematical
structure is well understood from an algebraic point of view. On the
other hand the models possess a formulation in terms of an action, and
thus more conventional techniques may be used. In addition, even though
rather special, a very large class of conformal field theories can be
constructed from WZW models, using the coset construction \cite{GKO}.
\smallskip
\noindent Most of the work which has been done on WZW models has only taken
into account the local structure of the underlying (target space)
group, ignoring global topological effects. In this paper we shall try
to understand some of these global issues. In particular we shall be
interested in WZW models of groups which are not simply-connected.
\noindent Whereas the algebraic approach to WZW models is more powerful for
local considerations, the formulation in terms of an action allows a
discussion of the global issues, and we shall thus take it as our
starting point. We shall explain how the theory can be quantised, and
show how the spectrum of the corresponding quantum theories can be
described algebraically. We shall then analyse how many inequivalent
quantisations exist and exhibit them explicitly. We find that the
various quantisations are parametrised by
$\mbox{Hom}(\pi_1(G),\mbox{Hom}(\pi_1(G),U(1)))$, where $\pi_1(G)$ is
the fundamental group of the group $G$ under consideration.
\noindent Having given the various quantisations explicitly, we shall study
some of their properties in detail. In particular, we shall analyse
the behaviour of the fields under monodromy, {\it i.e.} the analytic
continuation of one field about another one, and we shall show that,
for general quantisations, the operator product transforms with
respect to a (non-trivial) one-dimensional representation. This
implies in particular that the amplitudes are only defined on some
covering space. The appearance of an `anyonic' representation is quite
generic for two-dimensional quantum field theories (see {\it e.g.}
\cite{FM,Fre}), and it suggests that the theory is genuinely
braided. In a string theory inspired context, such theories have
traditionally been excluded; however, from the point of view of
Euclidean conformal field theory, this restriction does not seem to be
justified.
\noindent The class of theories for which all operator products are
invariant under monodromy, the {\it monodromy invariant theories}, are
of special interest, not least from the traditional point of view of
Euclidean conformal field theory. We shall show that, depending on the
structure of the fundamental group of $G$, one, two or eight of the
different quantisations are monodromy invariant. In more detail, for
$\pi_1(G)\cong {\Bbb Z}_N$, there exists one monodromy invariant theory,
unless $N$ is even, in which case there are two. One of the monodromy
invariant theories is the known modular invariant theory
\cite{Ber,FGK2}, and the other (for even $N$) is not modular
invariant. For $\pi_1(G)\cong {\Bbb Z}_2\times{\Bbb Z}_2$, there are eight
monodromy invariant quantisations, two of which are the modular
invariant theories of Felder {\it et.al.} \cite{FGK2}, and the other
six are not modular invariant.
\noindent The essential calculational tool for the analysis of the
monodromy is a formula for the adjoint action of non-contractible
loops in $LG$ on the generator $L_0$ which we derive. This
formula may have some interest in its own right as the action of the
non-contractible loops in $LG$ on ${\Bbb R}\oplus l{\frak g}\oplus{\Bbb R}$, the Lie
algebra of $\Pi\triangleright {\tilde L}{\widetilde G}$ \cite{PS}, a priori involves a
choice; here this choice is fixed by identifying the generator of the
rigid rotations, $L_0$, with the Sugawara expression which is quadratic
in the Lie algebra of ${\tilde L} {\widetilde G}$, $\hat{{\frak g}}\cong l{\frak g}\oplus{\Bbb R}$.
\medskip
\noindent The paper is organised as follows. We start in section~2 by
studying the quantisation of the theory defined by an action, and
explain how the different quantisations arise, giving explicit
formulae for the spectrum of all possible quantisations. In section~3,
we determine the adjoint action of non-contractible loops in $LG$ on
the generator $L_0$, and show that all quantisations transform with
respect to a one-dimensional representation under monodromy. We then
analyse in section~4 which quantisations are monodromy invariant. In
section~5, some examples for the additional (monodromy invariant)
theories are exhibited in detail, and section~6 contains a few
concluding remarks. In the appendix, we recall some of the less widely
known facts about the affine Weyl group, and give a new geometrical
proof for an old observation of Olive and Turok \cite{OT} about the
symmetries of the affine Dynkin diagram.
\section{The different quantisations}
\renewcommand{\theequation}{2.\arabic{equation}}
\setcounter{equation}{0}
We want to consider the theory defined by the WZW action
\cite{Wit,FGK1,FGK2}
\begin{equation}
\label{action}
{\cal S}[g] = -\frac{k}{4 \pi} \int_{{\cal M}}
\left\langle g^{-1} \partial g, g^{-1} \bar{\partial} g \right\rangle
- \frac{k}{24 \pi} \int_{{\cal B}}
\left\langle {\tilde g}^{-1} d {\tilde g},
[{\tilde g}^{-1} d {\tilde g}, {\tilde g}^{-1} d {\tilde g}] \right\rangle \,.
\end{equation}
Here the field $g:{\cal M} \rightarrow G$ takes values in a simple,
connected, compact group $G$, and ${\tilde g}$ is an extension of $g$ to
${\cal B}$ where $\partial {\cal B}={\cal M}$. ${\cal M}$ is the two-dimensional space-time,
and we take space to be compactified so that ${\cal M}= S^{1} \times {\rm I\!R}$.
$\langle .,.\rangle$ is the Killing form on the Lie algebra ${\frak g}$ of
$G$, normalised so that the longest roots of the algebra have length
square equal to $2$.
\noindent We shall mainly be interested in the case where $G$ is not
simply-connected. $G$ can then be written as
\begin{equation}
G = {\widetilde G} / C \,,
\end{equation}
where ${\widetilde G}$ is the universal covering group of $G$, and $C$ is a
subgroup of the centre ${\cal Z}$ of ${\widetilde G}$. The centre is isomorphic
to ${\Bbb Z}_N$ (for some $N$) for all simply-connected simple compact
groups, with the exception of $D_{2 n} \cong SO(4n)$ for which the
centre is ${\Bbb Z}_2\times{\Bbb Z}_2$.
\noindent The second term in (\ref{action}), the Wess-Zumino term, depends
on the choice of the extension ${\tilde g}$, and the consistency of the quantum
theory requires this ambiguity to be $2\pi {\Bbb Z}$. This
imposes quantisation conditions on $k$. For $C={\Bbb Z}_{N}$, the
quantisation condition is \cite{FGK2}
\begin{equation}
\label{quantis}
\frac{k N}{2} \langle c_1, c_2 \rangle \in {\Bbb Z} \,,
\end{equation}
for all $c_i\in C$, and for $C={\Bbb Z}_2\times{\Bbb Z}_2$,
the condition is
\begin{equation}
k \langle c_1, c_1 \rangle, \hspace*{0.5cm}
k \langle c_2, c_2 \rangle, \hspace*{0.5cm}
k \langle c_1, c_2 \rangle \in {\Bbb Z} \,,
\end{equation}
where $c_1=(1,0)$ and $c_2=(0,1)$. Here we have identified (as we
shall do from now on) $\hat{c}\in C\subset {\cal Z}\subset {\widetilde G}$ with
$c$ in the quotient space of a Cartan subalgebra ${\frak h}$ by the coroot
lattice via $\hat{c}=exp(c)$. For $C={\Bbb Z}_2\times{\Bbb Z}_2$, the definition of
the Wess-Zumino term remains ambiguous and there are at least two
quantisations \cite{FGK2}.
\noindent Recall that the most general solution $g(x,t)$ of the (classical)
WZW-model is of the form
\begin{equation}
g(x,t) = u(x^+) \cdot v(x^-) \,,
\end{equation}
where $x^{\pm} = t \pm x$ and $u,v:{\rm I\!R} \rightarrow G$. The solution
$g$ has to be periodic in $x\mapsto x+2 \pi$, and this implies that
$u$ and $v$ have to satisfy
\begin{equation}
\label{monodromy}
u(z+2 \pi) = u(z) M, \hspace*{0.5cm} v(z+2 \pi) = M v(z) \,,
\end{equation}
where $M$ is some group element. The pair of functions $(u,v)$ does
not uniquely determine the solution $g$; the transformation
\begin{equation}
u \rightarrow u g_0 \hspace*{0.5cm} v \rightarrow g_0^{-1} v \,,
\end{equation}
under which
\begin{equation}
M \rightarrow g_0^{-1} M g_0
\end{equation}
leaves $g$ invariant. We can use this freedom to rotate $M$ into a
fixed maximal torus $H$ of $G$. This fixes the pair of functions
$(u,v)$ generically up to
\begin{equation}
\label{gauge}
u \rightarrow u n \hspace*{0.5cm} v \rightarrow n^{-1} v \,,
\end{equation}
where $n\in N(H)$, the normaliser of $H$, under which $M\rightarrow
n^{-1} M n$ \cite{MPHOS}. (If $M$ is not regular, {\it i.e.} if $M$
belongs to more than one maximal torus, the residual symmetry is even
larger.) The action of $n$ on $M$ has kernel $H$; the quotient
${\cal W}=N(H)/H$ is the Weyl group of $G$.
\medskip
\noindent The configuration space of the WZW model is the space of
functions ${\cal M}\rightarrow G$. We could now attempt to describe the phase
space of the system in terms of coordinates (on this configuration
space) and their conjugate momenta, but this would seem to be rather
difficult. We shall therefore use a different description which was
already employed in \cite{MPHOS} and goes back to \cite{CWZ}. In this
approach, the phase space is regarded as the manifold of solutions of
the equations of motion of the field theory. As in \cite{MPHOS}, this
space can be regarded as a symplectic quotient of a larger space in
which we relax the constraint that the monodromy of $u$ and $v$ should
be the same. That is, we introduce $\nu_L, \nu_R \in {\frak h}$, the Lie
algebra of the maximal torus $H$, and write the left and right
monodromy, respectively, as
\begin{equation}
M_L= exp(2 \pi \nu_L) \hspace{0.5cm} M_R= exp(2 \pi \nu_R)\,.
\end{equation}
We can then move all the non-trivial homotopic information about
$u$ and $v$ into $\nu_L$ and $\nu_R$, respectively, {\it i.e.} we can
write
\begin{equation}
u(x^+) = {\tilde u}(x^+) exp(\nu_L x^+)
\end{equation}
and
\begin{equation} v(x^-) = exp(\nu_R x^-) {\tilde v}(x^-)\,,
\end{equation}
where ${\tilde u},{\tilde v}\in L{\widetilde G}$, the loop group of the simply connected
covering group ${\widetilde G}$ of $G$. We can restrict $\nu_L$ and $\nu_R$,
without loss of generality, to lie in a fixed alcove (a chamber of the
Stiefel diagram) of ${\frak g}$ \cite{BtD}. Then the phase space is the
quotient of the submanifold $exp(2 \pi \nu_L) = exp(2 \pi \nu_R)$ (as
a relation in $G$) by the action of $(H\times C)$,
\begin{equation}
\label{phase}
\cup_{c\in C} \left( \left. ({\tilde u},\nu_L,\nu_R,{\tilde v})
\right|_{\nu_L=c\cdot \nu_R} \right) / (H\times C) \,,
\end{equation}
where $H$ acts on ${\tilde u}$ and ${\tilde v}$ as in (\ref{gauge}) and leaves
$\nu_L$ and $\nu_R$ invariant, and $c\in C$ maps
$({\tilde u},\nu_L,\nu_R,{\tilde v})$ to $({\tilde u},\nu_L,\nu_R,{\tilde v}\, c)$. To describe
the action of $c\in C$ on $\nu_R$ we regard (as before) $C\cong
\pi_1(G)$ as the quotient space of the lattice of integral elements of
${\frak h}$, {\it i.e.} those that are mapped to $1$ in $G$, by the coroot
lattice; this quotient space acts naturally on a fixed alcove by
translation.
\medskip
\noindent To quantise the theory we should now find a subalgebra of
functions on phase space which can be consistently defined as
operators in the quantum theory, replacing Poisson brackets by
commutators. Unfortunately, the phase space is rather complicated (in
particular it is not a vector space), and it is therefore very
difficult to find such a subalgebra explicitly. On the other hand, we
might argue that the subalgebra should contain the analogues of the
position and momentum function, and that therefore the phase space
itself (regarded as a subspace of the space of distributions on phase
space) should be contained in the closure of this subalgebra. Then,
the quantum states should form a representation of (\ref{phase}), and
thus of
\begin{equation}
\label{phasesectors}
\left( \cup_{c\in C} \left. ({\tilde u},\nu_L,\nu_R,{\tilde v})
\right|_{\nu_L=c\cdot \nu_R} \right) \,,
\end{equation}
which is covariant under the action of $(H\times C)$. (This
description resembles strongly the formulation in \cite{MG4}.) For
$G={\widetilde G}$, $C$ is the trivial group, and (\ref{phase}) consists of one
component only. In this case the configuration space (and the phase
space) is simply connected, and there should exist only one
quantisation in which the quantum states form a representation of
(\ref{phasesectors}) which is invariant under $H$. This should force
the representations of ${\tilde u}$ and ${\tilde v}$ to be conjugate, and we expect
therefore that the spectrum only contains states in the diagonal
theory. Indeed, it is known \cite{FGK2} that the theory for ${\widetilde G}$ is
\begin{equation}
\label{diagonal}
{\cal H}_{{\widetilde G}} = \sum_{j} {\cal H}_j \otimes {\cal H}_{\overline{j}} \,,
\end{equation}
where the sum extends over all unitary positive energy representations
of ${\tilde L}{\widetilde G}$, the central extension of the loop group $L{\widetilde G}$
\cite{PS}, and ${\cal H}_{\overline{j}}$ denotes the conjugate representation to ${\cal H}_j$.
\noindent In the general case, the spectrum of the theory corresponding to
$G={\widetilde G}/C$ contains also states which are representations of the other
components of (\ref{phasesectors}). The additional states of the
quantum theory should be obtained by the action of the non-trivial
loops of $G$ (which are labelled by $c\in C$) on one of the two
representations in the tensor products of (\ref{diagonal}). Recall
from Pressley and Segal \cite{PS}, that the action (by conjugation) of
the non-trivial loops of $G$ on ${\tilde L}{\widetilde G}$ is well-defined. This action
does not preserve the set of positive roots, but, as has been shown in
\cite{FGK2} (see also the appendix), there exists a unique element in
the affine Weyl group ${\overline{\cal W}}({\frak g})$, so that the composition with this
affine Weyl group element preserves the positive roots. The
non-trivial loops of $G$ therefore induce (outer) automorphisms of
${\tilde L}{\widetilde G}$, and thus map in general a representations of ${\tilde L}{\widetilde G}$,
${\cal H}_j$, into a different representation which we denote by
${\cal H}_{c(j)}$. (The map $c(j)$ has been calculated for all simple groups
in \cite{FGK2}.)
\noindent Each component of the configuration space (and also of the phase
space (\ref{phase})) is not simply connected, the fundamental group
being isomorphic to $C$. We therefore expect that there should be
different quantisations of the classical theory corresponding to
different monodromies with respect to $C$. \footnote{As $C$ is
abelian, only one-dimensional representations of the fundamental group
appear.} To classify the different quantisations, recall that for any
subgroup $C$ of the centre of ${\widetilde G}$, the irreducible representations
of ${\widetilde G}$ (and thus also of ${\tilde L}{\widetilde G}$) fall into equivalence classes
which are characterised by the induced (one-dimensional)
representation of $C$. The tensor product of two irreducible
representations is in the equivalence class corresponding to the
product representation. For each sector, labelled by $c\in C$, let
$R_c$ denote the equivalence class of representations of ${\tilde L}{\widetilde G}$
which corresponds to the representation $R_c: C \rightarrow U(1)$. The
possible quantisations are then of the form
\begin{equation}
\label{gentheo}
{\cal H}_G^{R} : = \sum_{c\in C} \sum_{j\in R_c} {\cal H}_j \otimes {\cal H}_{c(\overline{j})}\,,
\end{equation}
where $R_c$ has to satisfy
\begin{equation}
\label{Rconsis}
R_{c_1} R_{c_2} = R_{c_1 + c_2} \,,
\end{equation}
so that the theory is closed under operator products. This means
that $R$ must be a homomorphism from $C\cong \pi_1(G)$ to
$\mbox{Hom}(\pi_1(G),U(1))$. For $C\cong {\Bbb Z}_N$,
\begin{equation}
\mbox{Hom}({\Bbb Z}_N,\mbox{Hom}({\Bbb Z}_N,U(1))) \cong
\mbox{Hom}({\Bbb Z}_N,U(1))\,,
\end{equation}
as each $R\in \mbox{Hom}({\Bbb Z}_N,\mbox{Hom}({\Bbb Z}_N,U(1)))$ is already
uniquely determined by $R_{c_1}$, the representation corresponding to
the generator $c_1$ of ${\Bbb Z}_N$. On the other hand, for
$C\cong{\Bbb Z}_2\times{\Bbb Z}_2$,
\begin{equation}
\mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,\mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,U(1))) \cong
\mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,U(1))\times
\mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,U(1)) \,,
\end{equation}
as every $R\in \mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,
\mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,U(1)))$ is
uniquely determined by $R_{(1,0)}$ and $R_{(0,1)}$.
\noindent Thus, there should exist different quantisations of the
WZW model corresponding to $G={\widetilde G}/C$, which are labelled by
$\mbox{Hom}(C,U(1))$ (for $C\cong {\Bbb Z}_N$) and by
$\mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,U(1))\times
\mbox{Hom}({\Bbb Z}_2\times{\Bbb Z}_2,U(1))$ (for $C\cong
{\Bbb Z}_2\times{\Bbb Z}_2$).
\noindent In general, these theories are not invariant under the monodromy
corresponding to the analytic continuation of one field
about another one. However, as we shall show in the next section, the
operator product transforms covariantly with respect to a
one-dimensional representation. The appearance of an `anyonic'
representation, describing the monodromy, indicates that the
corresponding theories are genuinely braided. The appearance of the
braid group is quite generic for two-dimensional quantum field
theories \cite{FM,Fre,FGR}. On the other hand, such theories have
traditionally been excluded in Euclidean conformal field theory.
\noindent The theories are also not modular invariant in general. However,
as we shall show, the modular invariant theory of Felder, Gawedzki and
Kupiainen \cite{FGK2} is one of the possible quantisations. It
corresponds to a quantisation which is invariant under monodromy, but
it is not characterised by this property alone. In fact, we shall show
that for $C={\Bbb Z}_N$ with $N$ even, there exists another monodromy
invariant theory (which is then not modular invariant), and for
$C={\Bbb Z}_2 \times {\Bbb Z}_2$, there exist in addition three monodromy
invariant quantisations for each of the two different modular
invariant solutions.
\noindent We should also mention that the above analysis resembles somewhat
the treatment in \cite{SY}, where simple currents were used to
construct modular invariant partition functions via an orbifold
construction.
\smallskip
\noindent From a general point of view, regarding the quantum states as
wave-functions on configuration space, one would expect that the
different quantisations correspond to the different choices for the
`Aharanov-Bohm' phases in each component of the configuration
space. If we insist that the quantum theory should be symmetric under
the full loop group (which acts naturally on the configuration space),
then the choice for the `Aharanov-Bohm' phases in the identity
component fixes the phases in all other components. We would thus
expect that the different quantisations are parametrised by
$\mbox{Hom}(\pi_1(G),U(1))$ as explained in \cite{MPHOS}. However,
since we are interested in a quantum field theory where states and
fields are in one-to-one correspondence, we have the additional
constraint that the theory should be closed under the operator
product. It is natural to believe that the Aharanov-Bohm phases
multiply under the operator product, and then the closure condition
implies that the phases of the identity component have to be
trivial. We therefore expect that there exists only one quantum field
theory which is symmetric under the full loop group, the theory with
trivial Aharanov-Bohm phases. This theory, however, is not modular
invariant in general as the modular invariant theories do not always
satisfy $R_c=id$ for all $c\in C$ \cite{FGK2}.
\noindent The above theories (\ref{gentheo}) are only symmetric under the
identity component of the loop group, and thus, a priori, the phases
in the different components are unrelated. The possible theories are
then selected by the condition that they are closed under operator
products. As the phases multiply, this implies that the different
quantisations are parametrised by
$\mbox{Hom}(\pi_1(G),\mbox{Hom}(\pi_1(G),U(1)))$. Depending on the
structure of the fundamental group one or two of these theories are
the modular invariant theories of \cite{FGK2}. These theories were
selected in \cite{FGK2} by the specific choice \cite[eq.\ (4.9)]{FGK2},
relating the Aharanov-Bohm phases in the different components.
\section{Monodromy covariance}
\renewcommand{\theequation}{3.\arabic{equation}}
\setcounter{equation}{0}
The key step in the analysis of the monodromy is the calculation of
the transformation of the spectrum of $L_0$ under conjugation by a
loop corresponding to $c\in C$. A priori, as described in Pressley and
Segal \cite{PS}, the action (by conjugation) of the non-trivial loops
of $G$ on ${\tilde L}{\widetilde G}$ is well-defined, and thus induces a well-defined
action on the (untwisted) Kac-Moody algebra $\hat{{\frak g}} \cong
l{\frak g}\oplus{\Bbb R}$, the Lie algebra of ${\tilde L}{\widetilde G}$. On the other hand, a
priori a choice has to be made for the definition of the conjugation
on ${\Bbb R}\oplus l{\frak g} \oplus{\Bbb R}$, the Lie algebra of the extension
$\Pi\triangleright{\tilde L}{\widetilde G}$ of ${\tilde L}{\widetilde G}$ by the rigid rotations $\Pi$
whose generator is $L_0$ (see section~4.9 of \cite{PS} and the
appendix). This ambiguity can be removed by identifying $L_0$ with the
Sugawara expression which is quadratic in $\hat{{\frak g}}$.
\noindent The Kac-Moody algebra $\hat{{\frak g}}$ can be described in a modified
Cartan-Weyl basis as follows \cite{GO}:
\begin{equation}
\begin{array}{ccl}
{\displaystyle [H^i_m, H^j_n]} & = &
{\displaystyle k m \delta^{ij} \delta_{m,-n} \vspace*{0.3cm} } \\
{\displaystyle [H^i_m, E^{\alpha}_n]} & = &
{\displaystyle \alpha^i E^{\alpha}_{m+n} \vspace*{0.3cm} } \\
{\displaystyle [E^{\alpha}_m, E^{\beta}_n]} & = &
\left\{
\begin{array}{ll}
{\displaystyle \varepsilon(\alpha,\beta) E^{\alpha+\beta}_{m+n}
\vspace*{0.2cm} } &
\mbox{if $\alpha+\beta$ is a root, \vspace*{0.2cm} } \\
{\displaystyle \frac{2}{\alpha^2} \left( \alpha \cdot H_{m+n} + k m
\delta_{m,-n} \right)} & \mbox{if $\alpha=-\beta$ ,\vspace*{0.2cm} } \\
{\displaystyle 0} & \mbox{otherwise.\vspace*{0.3cm} }
\end{array} \right. \\
{}[k,E^{\alpha}_n] & = & [k,H^i_n] = 0\,.
\end{array}
\end{equation}
Here $i=1, \ldots, r=\mbox{rank}~{\frak g}$, $\alpha$ labels the positive roots
$R^+$ of ${\frak g}$, and the horizontal subalgebra (with $n=0$) is isomorphic
to ${\frak g}$.
\smallskip
\noindent Upon conjugation with the loop $\theta\in [0,2 \pi ] \mapsto
exp(\theta c)$, where $exp(2 \pi c)\in{\cal Z}$, the centre of ${\widetilde G}$, the
generators of the Kac-Moody algebra transform as \cite{PS}
\begin{equation}
H^i_m \mapsto H^i_m - \delta_{m,0} k \langle c, H^i \rangle \,,
\end{equation}
\begin{equation}
E^{\alpha}_m \mapsto E^{\alpha}_{m - \langle \alpha,c \rangle} \,,
\end{equation}
\begin{equation}
k \mapsto k \,,
\end{equation}
where $\langle .,. \rangle$ denotes the Killing-form of the horizontal
subalgebra.
\medskip
\noindent The Sugawara expression for $L_0$ is given as \cite{GO}
$$
L_0 = \frac{1}{\beta} \left[
\sum_{i=1}^{r} H^i_0 H^i_0 +
\sum_{\alpha\in R^+} \left( E^{\alpha}_0 E^{-\alpha}_0 +
E^{- \alpha}_0 E^{\alpha}_0 \right) \right. \hspace*{4cm}
$$
\begin{equation}
\hspace*{3cm}
\left. + 2 \sum_{n=1}^{\infty}
\left( \sum_{i=1}^{r} H^i_{-n} H^i_n +
\sum_{\alpha\in R^+} \left( E^{\alpha}_{-n} E^{-\alpha}_n
+ E^{-\alpha}_{-n} E^{\alpha}_n \right) \right) \right] \,,
\end{equation}
where $\beta=2 k + Q_{\psi}$, and $Q_{\psi}$ is the quadratic Casimir
in the adjoint representation (with highest weight $\psi$, where
$\psi$ is the highest root). Upon conjugation with the loop
$\theta\in [0,2 \pi ] \mapsto exp(\theta c)$ $L_0$ becomes
\begin{eqnarray}
L_0^{\prime} & = & \frac{1}{\beta} \left[
\sum_{i=1}^{r} \left( H^i_0 H^i_0 - 2 k \langle c, H^i \rangle H^i_0
+ k^2 \langle c, H^i \rangle \langle H^i,c \rangle \right)
\right. \nonumber \\
& & + \sum_{\alpha\in R^+} \left( E^{\alpha}_{- \langle \alpha, c\rangle}
E^{-\alpha}_{\langle \alpha, c \rangle} +
E^{-\alpha}_{\langle \alpha, c \rangle}
E^{\alpha}_{-\langle \alpha, c \rangle} \right) \nonumber \\
& & \left.
+ 2 \sum_{n=1}^{\infty} \left( \sum_{i=1}^{r} H^i_{-n} H^i_n
+ \sum_{\alpha\in R^+} \left( E^{\alpha}_{-n- \langle \alpha, c \rangle}
E^{-\alpha}_{n + \langle \alpha, c \rangle}
+ E^{-\alpha}_{-n + \langle \alpha, c \rangle}
E^{\alpha}_{n - \langle \alpha, c \rangle}\right) \right) \right] \,.
\end{eqnarray}
We have the identities
\begin{equation}
- \sum_{i=1}^{r} 2\, k \,\langle c, H^i \rangle H^i_0 \; | \lambda \rangle =
- 2 \, k \, \langle c,\lambda \rangle \; | \lambda \rangle \,,
\end{equation}
where $\lambda$ is the weight of the state $|\lambda\rangle$ on which
$L_0^{\prime}$ is evaluated, and
\begin{equation}
\sum_{i=1}^{r} k^2 \langle c, H^i \rangle \langle H^i,c \rangle =
k^2 \langle c,c \rangle \,.
\end{equation}
Furthermore, we can choose (without loss of generality) the set of
positive roots, $R^+$, such that $\langle \alpha, c \rangle \geq 0$ for
all $\alpha\in R^+$. Then we can rewrite $L_0^{\prime} | \lambda \rangle$
as
$$
L_0 | \lambda \rangle
+ \frac{1}{\beta}\Biggl[- 2 k \langle c,\lambda \rangle
+ k^2 \langle c,c \rangle \Biggr.
\hspace*{10.5cm}
\vspace*{-0.6cm}
$$
\begin{eqnarray}
\hspace*{0.1cm}
& & \left.+ \sum_{\alpha\in R^+}
\left( 2 \sum_{n=1}^{\langle \alpha,c\rangle - 1}
[ E^{-\alpha}_{-n + \langle \alpha,c\rangle} ,
E^{\alpha}_{n - \langle \alpha,c\rangle}]
+ [E^{-\alpha}_0, E^{\alpha}_0 ]
+ [ E^{-\alpha}_{\langle \alpha,c\rangle} ,
E^{\alpha}_{- \langle \alpha,c\rangle}] \right) | \lambda \rangle
\right] \nonumber \\
& = & L_0 | \lambda \rangle
+ \frac{1}{\beta}\left[- 2 k \langle c,\lambda \rangle
+ k^2 \langle c,c \rangle
+ 2 \sum_{\alpha\in R^+} \left( \sum_{n=0}^{\langle \alpha,c\rangle -1}
[ E^{-\alpha}_{-n + \langle \alpha,c\rangle} ,
E^{\alpha}_{n - \langle \alpha,c\rangle}]
- \frac{k}{\alpha^2} \langle \alpha,c\rangle \right) | \lambda \rangle
\right] \nonumber \\
& = & L_0 | \lambda \rangle
+ \frac{1}{\beta}\Biggl[- 2 k \langle c,\lambda \rangle
+ k^2 \langle c,c \rangle \Biggr. \nonumber \\
& & \left. + \sum_{\alpha\in R^+} \left(
\sum_{n=0}^{\langle \alpha,c\rangle -1}
\left\{ - \frac{4}{\alpha^2} \langle \alpha, \lambda \rangle
+ \frac{4}{\alpha^2} k (-n + \langle \alpha, c \rangle) \right\}
- \frac{2 k}{\alpha^2} \langle \alpha, c \rangle \right) \right] \nonumber \\
& = & L_0 | \lambda \rangle
+ \frac{1}{\beta}\left[- 2 k \langle c,\lambda \rangle
+ k^2 \langle c,c \rangle
+ \sum_{\alpha\in R^+} \left\{
- \frac{4}{\alpha^2} \langle \alpha, \lambda \rangle
\langle \alpha, c \rangle
+ \frac{4}{\alpha^2} k \left( \sum_{l=1}^{\langle \alpha,c\rangle} l\right)
- \frac{2 k}{\alpha^2} \langle \alpha, c \rangle \right) \right] \nonumber \\
& = & L_0 | \lambda \rangle
+ \frac{1}{\beta}\left[- 2 k \langle c,\lambda \rangle
+ k^2 \langle c,c \rangle
+ \sum_{\alpha\in R^+} \left\{
- \frac{4}{\alpha^2} \langle \alpha, \lambda \rangle
\langle \alpha, c \rangle
+ \frac{2 k}{\alpha^2} \langle \alpha,c\rangle \langle c, \alpha\rangle
\right\} \right]\,. \nonumber \\
\end{eqnarray}
Finally, we can use the identity (see {\it e.g.} \cite{GOS})
\begin{equation}
\sum_{\alpha\in R^+} \frac{4}{\alpha^2}\;\; |\alpha \rangle \langle
\alpha| = Q_{\psi} {\mathchoice {{\rm 1\mskip-4mu l}} {{\rm 1\mskip-4mu l}_r \,,
\end{equation}
where ${\mathchoice {{\rm 1\mskip-4mu l}} {{\rm 1\mskip-4mu l}_r$ is the unit matrix in the space of rank~${\frak g}$
matrices, to conclude that
\begin{equation}
\label{difference}
L_0^{\prime} \; | \lambda \rangle =
L_0 \; | \lambda \rangle - \langle c,\lambda \rangle \; | \lambda \rangle
+ \frac{1}{2} \, k \, \langle c,c \rangle \; | \lambda \rangle \,,
\end{equation}
as $\beta = 2 k + Q_{\psi}$.
\medskip
\noindent If $c$ is a coroot, the conjugation corresponds to a
transformation in the affine Weyl group ${\overline{\cal W}}({\frak g})$, and the induced
transformation on weights coincides with the formula given in
\cite{GO}. We should also mention that essentially this formula has
been derived in \cite{FGV} for ${\frak g}=a_n$, and, in a different context,
in \cite{Halpern}.
\bigskip
\noindent We are now in the position to analyse the monodromy properties of
the different quantisations (\ref{gentheo}). Recall that for any
weight $\overline{\lambda}$ in a subrepresentation of the tensor product of
two representations with highest weight $\overline{\lambda}_1$ and
$\overline{\lambda}_2$, respectively, and for any $exp(2\pi c)\in{\cal Z}$, we have
\begin{equation}
\langle c, \overline{\lambda} \rangle = \langle c, \overline{\lambda}_1 \rangle
+ \langle c, \overline{\lambda}_2 \rangle \;\; (\mbox{mod $1$}) \,.
\end{equation}
Furthermore, the product of two states $(\lambda_1,c_1(\overline{\lambda}_1))$
and $(\lambda_2,c_2(\overline{\lambda}_2))$ in (\ref{gentheo}) is in the
sector $(\lambda,(c_1 + c_2)(\overline{\lambda}))$. As the $L_0$ spectrum is
the same for a representation and its conjugate, upon rotating the
field corresponding to $(\lambda_1,c_1(\overline{\lambda}_1))$ by $2 \pi$
about $(\lambda_2,c_2(\overline{\lambda}_2))$, the three-point function
changes by
\begin{equation}
\label{braid}
R(\lambda_1, c_1; \lambda_2, c_2) =
exp \Bigl\{ 2 \pi \Bigl( - \langle \overline{\lambda}_1, c_2 \rangle
- \langle \overline{\lambda}_2, c_1 \rangle + k \langle c_1, c_2 \rangle \Bigr)
\Bigr\} \,,
\end{equation}
which is independent of $\lambda$. This demonstrates that the fields
of (\ref{gentheo}) are covariant under monodromy with respect to a
one-dimensional representation which is given by (\ref{braid}).
\section{Additional monodromy invariant theories}
\renewcommand{\theequation}{4.\arabic{equation}}
\setcounter{equation}{0}
We want to analyse now, how many monodromy invariant quantisations
exist. Recall from section~2 that the different quantisations are of
the form
\begin{equation}
\label{localtheo}
{\cal H}_G^R := \sum_{c\in C} \sum_{j\in R_c} {\cal H}_j \otimes {\cal H}_{c(\overline{j})} \,,
\end{equation}
where $G={\widetilde G}/C$ and $R_c$ is an equivalence class of positive energy
representations of $\hat{{\frak g}}$ corresponding to a representation of
$C$ (which we also denote by $R_c: C \rightarrow U(1)$).
In order for the theory to be closed under composition, the
assignment of $R_c$ to $c$ must respect the group structure of $C$,
{\it i.e.} $R$ must be an element of $\mbox{Hom}(C,\mbox{Hom}(C,U(1)))$.
The quantisation is invariant under monodromy, if $R(\lambda_1,
c_1;\lambda_2, c_2)=1$ for all $(\lambda_i,c_i(\overline{\lambda}_i))$ in
(\ref{localtheo}). Let us consider the two different cases for the
structure of $C$ separately.
\subsection{The case $C={\Bbb Z}_N$.}
Let $c$ denote a cyclic generator of $C$, where the order of $c$ is
$N$. Because of the representation property of $R$, all
$R_{c^\prime}$ are uniquely determined, once $R_c$ is fixed. Let
$\lambda\in R_c$. By requiring monodromy invariance for the product of
$(\lambda,c)$ with itself, we find using (\ref{braid})
\begin{equation}
0 = 2 \left( - \langle \overline{\lambda},c\rangle + \frac{1}{2} k \langle c,c\rangle
\right) \;\; (\mbox{mod $1$})\,.
\end{equation}
As $C$ has order $N$, any possible weight $\overline{\lambda}$ has to satisfy
\begin{equation}
\label{constr}
N \langle \overline{\lambda},c\rangle = 0 \;\; (\mbox{mod $1$}) \,,
\end{equation}
and any linear functional, satisfying (\ref{constr}), corresponds to a
class of possible weights. The quantisation condition
\begin{equation}
\label{quanti}
\frac{1}{2} N k \langle c,c\rangle \in {\Bbb Z}
\end{equation}
thus guarantees that there exists $\overline{\lambda}$ such that
\begin{equation}
- \langle \overline{\lambda},c\rangle + \frac{1}{2} k \langle c,c\rangle =0 \;\;
(\mbox{mod $1$})\,.
\end{equation}
In this case, the difference of the $L_0$ eigenvalues of the
representation corresponding to $\lambda$ and to $c(\overline{\lambda})$ is
integral because of (\ref{difference}) and since a representation and
its conjugate have the same $L_0$ spectrum. The same is true
for all other sectors, since
\begin{equation}
- \langle m \overline{\lambda}, m c \rangle + \frac{1}{2} k \langle m c, m c \rangle
= m^2 \left( - \langle \overline{\lambda},c\rangle + \frac{1}{2} k \langle c,c\rangle
\right) = 0 \;\; (\mbox{mod $1$}) \,.
\end{equation}
Thus this solution corresponds to the unique modular invariant theory
of Felder {\it et.al.} \cite{FGK2}.
\medskip
\noindent Because of the consistency condition (\ref{constr}) and the
quantisation condition (\ref{quanti}), there exists a solution
$\overline{\lambda}_1$, satisfying
\begin{equation}
- \langle \overline{\lambda}_1 , c \rangle + \frac{1}{2} k \langle c, c \rangle =
\frac{1}{2}\;\; (\mbox{mod $1$})\,,
\end{equation}
only if $N$ is even. On the other hand, for even $N$, $\overline{\lambda}_1$
is a possible weight as it corresponds to a representation in the
equivalence class of the one-dimensional representation of the centre
\begin{equation}
R(w) = - e^{\pi i k \langle w,w \rangle} \,,
\end{equation}
where $w\in C$. To check that the corresponding theory is monodromy
invariant, we observe that
\begin{equation}
- \langle m \overline{\lambda}_1, n c \rangle
- \langle n \overline{\lambda}_1, m c \rangle
+ k \langle m c, n c \rangle
= m n \left( - 2 \langle \overline{\lambda}_1, c \rangle
+ k \langle c, c \rangle \right) \in {\Bbb Z} \,.
\end{equation}
Thus all sectors are relatively monodromy invariant. It is clear
that this additional monodromy invariant theory is not modular
invariant, as the partition function is not invariant under
$\tau \mapsto \tau + 1$.
\subsection{The case $C={\Bbb Z}_2\times{\Bbb Z}_2$.}
Let $c_1 = (1,0)$ and $c_2=(0,1)$ be the two generators of
$C={\Bbb Z}_2\times{\Bbb Z}_2$, and denote by $\lambda_1$ and $\lambda_2$ the
corresponding weights in (\ref{localtheo}). We have to check, case by
case, the conditions implied by monodromy invariance..
\begin{list}{(\roman{enumi})}{\usecounter{enumi}}
\item $(1,0) \times (1,0) = (0,0)$. The requirement is
\begin{equation}
- 2 \langle \overline{\lambda}_1, c_1 \rangle + k \langle c_1, c_1 \rangle =
0 \;\; (\mbox{mod $1$}) \,.
\end{equation}
As $2c_1$ is a coroot, $2 \langle \overline{\lambda}_1, c_1 \rangle \in{\Bbb Z}$, and
the condition is
\begin{equation}
k \langle c_1, c_1 \rangle \in {\Bbb Z} \,,
\end{equation}
which is one of the quantisation conditions.
\item $(0,1) \times (0,1) = (0,0)$. An identical reasoning gives
\begin{equation}
k \langle c_2, c_2 \rangle \in {\Bbb Z} \,,
\end{equation}
which is again one of the quantisation conditions.
\item $(1,1) \times (1,1) = (0,0)$. Similarly we find
\begin{equation}
k \langle c_1+c_2, c_1+c_2 \rangle \in {\Bbb Z} \,,
\end{equation}
which follows from one of the quantisation conditions.
\item $(1,0) \times (0,1) = (1,1)$. Using (\ref{braid}), we find
\begin{equation}
- \langle \overline{\lambda}_1, c_2 \rangle - \langle \overline{\lambda}_2, c_1 \rangle
+ k \langle c_1, c_2 \rangle \in {\Bbb Z} \,.
\end{equation}
\end{list}
As explained in \cite{FGK2}, there are two modular invariant theories.
One is characterised by
\begin{equation}
\label{sol1}
\begin{array}{lcl}
{\displaystyle
\langle \overline{\lambda}_1,c_1 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_1, c_1
\rangle + \frac{n}{2} \;\; (\mbox{mod $1$})} \vspace*{0.2cm} \\
{\displaystyle \langle \overline{\lambda}_1,c_2 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_1, c_2
\rangle \;\;(\mbox{mod $1$})} \vspace*{0.2cm} \\
{\displaystyle \langle \overline{\lambda}_2,c_1 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_2, c_1
\rangle \;\;(\mbox{mod $1$})} \vspace*{0.2cm} \\
{\displaystyle \langle \overline{\lambda}_2,c_2 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_2, c_2
\rangle + \frac{m}{2} \;\;(\mbox{mod $1$}) \,,}
\end{array}
\end{equation}
where $m=n=0$, and the other is characterised by
\begin{equation}
\label{sol2}
\begin{array}{lcl}
{\displaystyle \langle \overline{\lambda}_1,c_1 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_1, c_1
\rangle + \frac{n}{2} \;\;(\mbox{mod $1$})} \vspace*{0.2cm} \\
{\displaystyle \langle \overline{\lambda}_1,c_2 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_1, c_2
\rangle +\frac{1}{2} \;\;(\mbox{mod $1$})} \vspace*{0.2cm} \\
{\displaystyle \langle \overline{\lambda}_2,c_1 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_2, c_1
\rangle +\frac{1}{2} \;\;(\mbox{mod $1$})} \vspace*{0.2cm} \\
{\displaystyle \langle \overline{\lambda}_2,c_2 \rangle} & = &
{\displaystyle \frac{k}{2} \langle c_2, c_2
\rangle +\frac{m}{2} \;\;(\mbox{mod $1$}) \,,}
\end{array}
\end{equation}
where, again, $m=n=0$.
\noindent To each of the two solutions, there exist additional monodromy
invariant solutions given by (\ref{sol1}) and (\ref{sol2}),
respectively, with $(n=1, m=0)$, $(n=0,m=1)$ and $(n=1,m=1)$. It is
immediate that they satisfy all constraints. It is clear that these
theories are not modular invariant, as the partition functions are not
invariant under $\tau \mapsto \tau + 1$.
\section{Examples}
\renewcommand{\theequation}{5.\arabic{equation}}
\setcounter{equation}{0}
In this section we shall give a few non-trivial examples, exhibiting
the additional (monodromy invariant) quantisations. The simplest
example occurs for $G=SO(3)$ and was already pointed out in
\cite{MPHOS}.
\subsection{$G=SO(3)$.}
The first homotopy group of $SO(3)$ is $\pi_1(SO(3))={\Bbb Z}_2$, and
the generator of $\pi_1(SO(3))$, written as an element of ${\frak h}$, is
$c=\sqrt{2}/2$. (Recall, that $E^\pm$ are $\pm\sqrt{2}$ in this
notation.) The quantisation condition requires $k$ to be even, as
$\langle c, c \rangle = 1/2$ (see (\ref{quantis})).
\noindent The outer automorphism corresponding to $c$ acts on
representations by mapping the representation with spin $j$ to
the one with spin $c(j)=k/2-j$. There are two cases to consider
\begin{list}{(\roman{enumi})}{\usecounter{enumi}}
\item $k=4n$. The modular invariant theory is given by
\begin{eqnarray}
{\cal H}_{0} & = &
\Bigl[ \Bigl([0]\otimes [0]\Bigr)
\oplus \Bigl[1]\otimes [1]\Bigr)
\oplus \Bigl[2]\otimes [2]\Bigr) \oplus \ldots \Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0]\otimes [k/2]\Bigr)
\oplus \Bigl([1]\otimes [k/2-1]\Bigr) \oplus \ldots \Bigr]
\nonumber \,.
\end{eqnarray}
\item $k=2n+2$. The modular invariant theory is given by
\begin{eqnarray}
{\cal H}_{1} & = &
\Bigl[ \Bigl([0]\otimes [0]\Bigr)
\oplus \Bigl([1]\otimes [1]\Bigr)
\oplus \Bigl([2]\otimes [2]\Bigr) \oplus \ldots \Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([1/2]\otimes [k/2-1/2]\Bigr)
\oplus \Bigl([3/2]\otimes [k/2-3/2]\Bigr) \oplus \ldots \Bigr]
\nonumber \,.
\end{eqnarray}
\end{list}
\noindent The additional quantisation which is also monodromy
invariant is for $k=4n$, ${\cal H}_1$, and for $k=2n+2$, ${\cal H}_0$. To check
that these theories are indeed monodromy invariant, we note that the
conformal weight of the lowest energy space of the representation
$[j]$ is
\begin{equation}
L_0([j]) = \frac{j(j+1)}{k+2} \,,
\end{equation}
and that
\begin{equation}
L_0([c(j)]) - L_0([j]) = \frac{k}{4} - j\,.
\end{equation}
This agrees with (\ref{difference}), as $\langle c, j \rangle = j$,
and the claimed monodromy invariance is easily verified.
\subsection{Quotient groups of $G=SU(4)$.}
\noindent The centre of $SU(4)$ is ${\cal Z}\cong {\Bbb Z}_4$, and the group of
central elements is generated by $c=\lambda_1$, the
fundamental weight corresponding to the first root (for the notation see
for example \cite{Gourdin}). The central
elements act (as outer automorphisms) on the representations
$[l,m,n]$, written in the Dynkin basis, as
\begin{equation}
\begin{array}{lcl}
c([l,m,n]) & = &
{\displaystyle [m,n,k-l-m-n]\,,} \\
c^2([l,m,n]) & = &
{\displaystyle [n,k-l-m-n,l]\,,} \\
c^3([l,m,n]) & = &
{\displaystyle [k-l-m-n,l,m]\,.}
\end{array}
\end{equation}
The conformal weight of the lowest energy space of the representation
$[l,m,n]$ is
\begin{equation}
\label{local2}
L_0([l,m,n]) = \Bigl( \frac{1}{4} (3 l^2 + 4 m^2 + 3 n^2 + 4 lm + 4mn +
2ln) + (3l + 4m + 3n) \Bigr) / (2(k+4)) \,.
\end{equation}
Using these explicit formulae, it is easy to check that
(\ref{difference}) holds.
\medskip
\noindent The centre contains the two different subgroups ${\Bbb Z}_2$ and
${\Bbb Z}_4$, and thus there are two different quotient groups whose
simply connected covering group is $SU(4)$. Let us discuss the two
cases in turn.
\begin{list}{(\roman{enumi})}{\usecounter{enumi}}
\item $G=SU(4)/ {\Bbb Z}_4$. The quantisation condition is $k\in 2{\Bbb Z}$,
as $\langle c, c \rangle= 3/4$ and $N=4$ (see (\ref{quantis})).
We have to consider the four cases $k=8p+2s$, where $p\in{\Bbb Z}$ and
$s=0,1,2,3$.
\begin{itemize}
\item $k=8p$. The modular invariant partition function is
\begin{eqnarray}
{\cal H}_0 & = &
\Bigl[ \Bigl([0,0,0]\otimes [0,0,0]\Bigr)
\oplus \Bigl([1,0,1]\otimes [1,0,1]\Bigr) \oplus \ldots
\Bigr] \nonumber \\
& & \oplus
\Bigl[ \Bigl([0,0,0]\otimes [0,0,k]\Bigr)
\oplus \Bigl([1,0,1]\otimes [0,1,k-2]\Bigr) \oplus \ldots
\Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,0,0]\otimes [0,k,0]\Bigr)
\oplus \Bigl([1,0,1]\otimes [1,k-2,1]\Bigr) \oplus \ldots
\Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,0,0]\otimes [k,0,0]\Bigr)
\oplus \Bigl([1,0,1]\otimes [k-2,1,0]\Bigr) \oplus \ldots
\Bigr] \,. \nonumber
\end{eqnarray}
\item $k=8p+2$. The modular invariant partition function is
\begin{eqnarray}
{\cal H}_1 & = &
\Bigl[ \Bigl([0,0,0]\otimes [0,0,0]\Bigr)
\oplus \Bigl([1,0,1]\otimes [1,0,1]\Bigr) \oplus \ldots \Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([1,0,0]\otimes [0,0,k-1]\Bigr)
\oplus \Bigl([0,1,1]\otimes [1,1,k-2]\Bigr) \oplus \ldots
\Bigr] \nonumber \\
& & \oplus
\Bigl[ \Bigl([0,1,0]\otimes [0,k-1,0]\Bigr)
\oplus \Bigl([1,1,1]\otimes [1,k-3,1]\Bigr) \oplus \ldots
\Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,0,1]\otimes [k-1,0,0]\Bigr)
\oplus \Bigl([1,1,0]\otimes [k-2,1,1]\Bigr) \oplus \ldots
\Bigr] \,. \nonumber
\end{eqnarray}
\item $k=8p+4$. The modular invariant partition function is
\begin{eqnarray}
{\cal H}_2 & = &
\Bigl[ \Bigl([0,0,0]\otimes [0,0,0]\Bigr)
\oplus \Bigl([1,0,1]\otimes [1,0,1]\Bigr) \oplus \ldots \Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,1,0]\otimes [1,0,k-1]\Bigr)
\oplus \Bigl([2,0,0]\otimes [0,0,k-2]\Bigr) \oplus \ldots
\Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,0,0]\otimes [0,k,0]\Bigr)
\oplus \Bigl([1,0,1]\otimes [1,k-2,1]\Bigr) \oplus \ldots
\Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,1,0]\otimes [k-1,0,1]\Bigr)
\oplus \Bigl([0,0,2]\otimes [k-2,0,0]\Bigr) \oplus \ldots
\Bigr] \,. \nonumber
\end{eqnarray}
\item $k=8p+6$. The modular invariant partition function is
\begin{eqnarray}
{\cal H}_3 & = &
\Bigl[ \Bigl([0,0,0]\otimes [0,0,0]\Bigr)
\oplus \Bigl([1,0,1]\otimes [1,0,1]\Bigr) \oplus \ldots \Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,0,1]\otimes [0,1,k-1]\Bigr)
\oplus \Bigl([1,1,0]\otimes [1,0,k-2]\Bigr) \oplus \ldots
\Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,1,0]\otimes [0,k-1,0]\Bigr)
\oplus \Bigl([1,1,1]\otimes [1,k-3,1]\Bigr) \oplus \ldots
\Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([1,0,0]\otimes [k-1,1,0]\Bigr)
\oplus \Bigl([0,1,1]\otimes [k-2,0,1]\Bigr) \oplus \ldots
\Bigr] \,. \nonumber
\end{eqnarray}
\end{itemize}
\noindent The different quantisations are for each even $k$,
${\cal H}_0, \ldots ,{\cal H}_3$. The additional monodromy invariant quantisation
is, for $k=8p$ ${\cal H}_2$, for $k=8p+2$ ${\cal H}_3$, for $k=8p+4$ ${\cal H}_0$, and
for $k=8p+6$ ${\cal H}_1$. Using (\ref{local2}), it is easy to see that
these theories are indeed monodromy invariant.
\item $G=SU(4)/ {\Bbb Z}_2$. The generator of ${\Bbb Z}_2$ is $d=2c$.
The quantisation condition is $k\in{\Bbb Z}$, as $\langle d,d \rangle = 4$
and $N=2$ (see (\ref{quantis})). We have to consider two cases, $k$
even and $k$ odd.
\begin{itemize}
\item For $k$ even, the modular invariant theory is given by
\begin{eqnarray}
{\cal H}_0 & = &
\Bigl[ \Bigl([0,0,0]\otimes [0,0,0]\Bigr)
\oplus \Bigl([0,1,0]\otimes [0,1,0]\Bigr) \oplus \ldots \Bigr]
\nonumber \\
& & \oplus
\Bigl[ \Bigl([0,0,0]\otimes [0,k,0]\Bigr)
\oplus \Bigl([0,1,0]\otimes [0,k-1,0]\Bigr) \oplus \ldots
\Bigr]\,. \nonumber
\end{eqnarray}
\item For odd $k$, the modular invariant theory is given by
\begin{eqnarray}
{\cal H}_1 & = &
\Bigl[ \Bigl([0,0,0]\otimes [0,0,0]\Bigr)
\oplus \Bigl([0,1,0]\otimes [0,1,0]\Bigr) \oplus \ldots
\Bigr] \nonumber \\
& & \oplus
\Bigl[ \Bigl([1,0,0]\otimes [0,k-1,1]\Bigr)
\oplus \Bigl([0,0,1]\otimes [1,k-1,0]\Bigr) \oplus \ldots \Bigr]
\,. \nonumber
\end{eqnarray}
\end{itemize}
\noindent Again, as before, the additional quantisation which is also
monodromy invariant is for $k$ even ${\cal H}_1$, and for $k$ odd ${\cal H}_0$.
As before, the monodromy invariance can be easily checked using
(\ref{local2}).
\end{list}
\subsection{$G=SO(8)/{\Bbb Z}_2\times {\Bbb Z}_2$.}
The centre of $G=SO(8)$ is ${\cal Z}\cong{\Bbb Z}_2\times {\Bbb Z}_2$, and it is
generated by $(1,0) = \lambda_1$, $(0,1) = \lambda_3$ and $(1,1) =
\lambda_4$, where $\lambda_i$ is the fundamental weight corresponding
to the $i$th root. The central elements act (as outer automorphisms)
on the representations $[r_1,r_2,r_3,r_4]$, written in the Dynkin
basis, as \begin{equation}
\begin{array}{lcl}
c_{(1,0)} ([r_1,r_2,r_3,r_4]) & = &
{\displaystyle [k-r_1-2 r_2-r_3-r_4,r_2,r_4,r_3]\,,} \\
c_{(0,1)} ([r_1,r_2,r_3,r_4]) & = &
{\displaystyle [r_4,r_2,k-r_1-2 r_2-r_3-r_4,r_1]\,,} \\
c_{(1,1)} ([r_1,r_2,r_3,r_4]) & = &
{\displaystyle [r_3,r_2,r_1,k-r_1-2 r_2-r_3-r_4]\,.}
\end{array}
\end{equation}
The conformal weight of the lowest energy space of the representation
$[r_1,r_2,r_3,r_4]$ is
$$
L_0([r_1,r_2,r_3,r_4]) = \Bigl( r_1^2 + 2 r_2^2 + r_3^2 + r_4^2
+ 2 r_2 (r_1 + r_3 + r_4) + r_1 r_3 + r_1 r_4 + r_3 r_4 \Bigr.
\hspace*{2cm}
$$
\begin{equation}
\hspace*{3cm} \Bigl.
+ 6 r_1 + 10 r_2 + 6 r_3 + 6 r_4 \Bigr) / \left(2 (k+6)\right)\,.
\end{equation}
It is easy to see that the formula for the difference of the
$L_0$ spectrum (\ref{difference}) holds.
\medskip
\noindent The quantisation condition is $k\in 2 {\Bbb Z}$, and for each allowed
$k$ there are two different modular invariant theories. They are
explicitly given as
\begin{equation}
\begin{array}{lcl}
{\cal H}_0 & = & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
[r_1,r_2,r_3,r_4] \Bigr) \right| r_{13}\;\mbox{even};
r_{14} \;\mbox{even}; r_{34} \;\mbox{even}
\Bigr]} \vspace*{0.2cm} \\
& & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
c_{(1,0)}([r_1,r_2,r_3,r_4]) \Bigr) \right|
r_{34}\; \mbox{even}; r_{13}+\frac{k}{2}\;\mbox{even};
r_{14}+\frac{k}{2}\;\mbox{even} \Bigr]} \vspace*{0.2cm} \\
& & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
c_{(0,1)}([r_1,r_2,r_3,r_4]) \Bigr) \right|
r_{14} \; \mbox{even}; r_{13}+\frac{k}{2}\;\mbox{even};
r_{34}+\frac{k}{2} \; \mbox{even} \Bigr]} \vspace*{0.2cm} \\
& & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
c_{(1,1)}([r_1,r_2,r_3,r_4]) \Bigr) \right|
r_{13} \; \mbox{even}; r_{14}+\frac{k}{2}\;\mbox{even};
r_{34}+\frac{k}{2} \;\mbox{even} \Bigr]\,,}
\end{array}
\end{equation}
and
\begin{equation}
\begin{array}{lcl}
{\cal H}_1 & = & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
[r_1,r_2,r_3,r_4] \Bigr) \right| r_{13}\;\mbox{even};
r_{14}\;\mbox{even}; r_{34} \; \mbox{even}
\Bigr]} \vspace*{0.2cm} \\
& & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
c_{(1,0)}([r_1,r_2,r_3,r_4]) \Bigr) \right|
r_{34}\; \mbox{even}; r_{13}+\frac{k}{2}\;\mbox{odd};
r_{14}+\frac{k}{2} \;\mbox{odd} \Bigr]} \vspace*{0.2cm} \\
& & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
c_{(0,1)}([r_1,r_2,r_3,r_4]) \Bigr) \right|
r_{14} \; \mbox{even}; r_{13}+\frac{k}{2}\;\mbox{odd};
r_{34}+\frac{k}{2} \;\mbox{odd} \Bigr]} \vspace*{0.2cm} \\
& & \oplus
{\displaystyle \Bigl[ \left. \Bigl( [r_1,r_2,r_3,r_4] \otimes
c_{(1,1)}([r_1,r_2,r_3,r_4]) \Bigr) \right|
r_{13} \; \mbox{even}; r_{14}+\frac{k}{2}\;\mbox{odd};
r_{34}+\frac{k}{2} \;\mbox{odd} \Bigr]\,,}
\end{array}
\end{equation}
where $r_{ij}=r_i + r_j$, and only those representations appear which
are positive energy, {\it i.e.} satisfy $r_1+2 r_2 + r_3 + r_4 \leq
k$. To shorten the notation, let us describe in the following the theories
by $4 \times 3$ matrices with entries $e$ (even) and $o$ (odd), where the
first row corresponds to $(r_{13}, r_{14}, r_{34})$ in the identity sector,
the second row corresponds to $(r_{34}, r_{13}+ k /2, r_{14} + k/2)$ in
the $c_{(1,0)}$ sector, and so on. For example
\begin{equation}
{\cal H}_0 = \left(\matrix{e & e & e \cr
e & e & e \cr
e & e & e \cr
e & e & e \cr}\right) \hspace*{2cm}
{\cal H}_1 = \left(\matrix{e & e & e \cr
e & o & o \cr
e & o & o \cr
e & o & o \cr}\right) \,.
\end{equation}
The additional monodromy invariant theories corresponding to
${\cal H}_0$ are then given as
\begin{equation}
{\cal H}_0^{(1)} = \left(\matrix{e & e & e \cr
o & o & e \cr
e & e & e \cr
o & e & o \cr}\right) \hspace*{2cm}
{\cal H}_0^{(2)} = \left(\matrix{e & e & e \cr
e & e & e \cr
o & o & e \cr
o & o & e \cr}\right) \hspace*{2cm}
{\cal H}_0^{(3)} = \left(\matrix{e & e & e \cr
o & o & e \cr
o & o & e \cr
e & o & o \cr}\right) \,,
\end{equation}
and the additional monodromy invariant theories
corresponding to ${\cal H}_1$ are
\begin{equation}
{\cal H}_1^{(1)} = \left(\matrix{e & e & e \cr
o & e & o \cr
e & o & o \cr
o & o & e \cr}\right) \hspace*{2cm}
{\cal H}_1^{(2)} = \left(\matrix{e & e & e \cr
e & o & o \cr
o & e & o \cr
o & e & o \cr}\right) \hspace*{2cm}
{\cal H}_1^{(3)} = \left(\matrix{e & e & e \cr
o & e & o \cr
o & e & o \cr
e & e & e \cr}\right)\,.
\end{equation}
{}From the formulae given above, it is easy to see that these theories
are indeed monodromy invariant. Furthermore, there are another $8$
quantisations which are not monodromy invariant
$$
{\cal H}_9 = \left(\matrix{e & e & e \cr
e & e & e \cr
e & o & o \cr
o & e & o \cr}\right) \hspace*{0.8cm}
{\cal H}_{10} = \left(\matrix{e & e & e \cr
e & e & e \cr
o & e & o \cr
e & o & o \cr}\right) \hspace*{0.8cm}
{\cal H}_{11} = \left(\matrix{e & e & e \cr
e & o & o \cr
e & e & e \cr
o & o & e \cr}\right) \hspace*{0.8cm}
{\cal H}_{12} = \left(\matrix{e & e & e \cr
e & o & o \cr
o & o & e \cr
e & e & e \cr}\right) \,,
$$
$$
{\cal H}_{13} = \left(\matrix{e & e & e \cr
o & e & o \cr
e & e & e \cr
e & o & o \cr}\right) \hspace*{0.8cm}
{\cal H}_{14} = \left(\matrix{e & e & e \cr
o & e & o \cr
o & o & e \cr
o & e & o \cr}\right) \hspace*{0.8cm}
{\cal H}_{15} = \left(\matrix{e & e & e \cr
o & o & e \cr
o & e & o \cr
o & o & e \cr}\right) \hspace*{0.8cm}
{\cal H}_{16} = \left(\matrix{e & e & e \cr
o & o & e \cr
e & o & o \cr
e & e & e \cr}\right) \,.
$$
\section{Conclusions}
We have analysed the different quantisations of the WZW model
corresponding to a in general non-simply connected group $G$, and we
have found that they are parametrised by
$\mbox{Hom}(\pi_1(G),\mbox{Hom}(\pi_1(G),U(1)))$. In general the
different quantisations are genuinely braided and neither monodromy
nor modular invariant. However, all the modular invariant theories of
Felder {\it et.al.} \cite{FGK2} are contained among them. Furthermore,
for $\pi_1(G)\cong {\Bbb Z}_N$ with $N$ even, there is another monodromy
invariant theory, and for $\pi_1(G)\cong {\Bbb Z}_2\times{\Bbb Z}_2$ there are
another $6$ monodromy invariant quantisations.
\noindent It might be hoped that the different quantisations of the WZW
model are characterised by the property to be {\it modular
covariant}. This would mean that it should be possible to define
amplitudes for these theories on arbitrary Riemann surfaces. These
amplitudes would not be invariant under the action of the modular
group. However, they would transform with respect to a
(one-dimensional) representation, and therefore the corresponding
probabilities would be invariant. (The amplitudes would be rather
similar to the amplitudes of a chiral theory, but here they would
correspond to the whole theory.) A first indication that this might be
the case is the fact that the amplitudes transform with respect to a
one-dimensional representation of the monodromy group, as shown in
section~3.
\smallskip
\noindent Braided conformal field theories similar to the ones discussed in
this paper have already appeared in \cite{Anni}.\footnote{I thank G.
Watts for drawing my attention to this reference.} There the conformal
limit (in the sector $w=0$) of the $\phi_{3,1}$ (integrable)
perturbation of ${\cal M}_{3,5}$ was found to be the theory
\begin{equation}
{\cal H}_{3,5}^{3,1} = \left({\cal H}_{1,1}\otimes{\cal H}_{1,1}\right)
\oplus \left( {\cal H}_{3,1}\otimes{\cal H}_{3,1}\right)
\oplus \left( {\cal H}_{2,1}\otimes{\cal H}_{3,1}\right)
\oplus \left( {\cal H}_{4,1}\otimes{\cal H}_{1,1}\right)
\end{equation}
which is genuinely braided. (Our notation follows, for example,
\cite{ID}.) In this context our analysis seems to suggest that the
conformal limit of the different massive perturbations of a conformal
field theory correspond to different quantisations and global
structures of the underlying conformal theory. It would be
interesting to check this conjecture by analysing the conformal limit
of massive perturbations of WZW models.
\noindent The analysis of the global properties of the WZW models might
also be relevant for a better understanding of the global issues of
(abelian) $T$-duality in WZW models \cite{Kir,AABL,AAL,RV}. In
particular, similar techniques might be used to analyse in which way
the duality transformation depends on the global topological
properties of the target space group. This is currently work in
progress.
\section*{Appendix}
|
\section{Introduction}
The discovery that $N>1$ supergravity theories lead to antigravity
is due to the work of the late J. Scherk \cite{Scherk,ScherkProc}.
In a recent paper we have revived the interest for the implications
of extended supergravity theories for antigravity \cite{belfar}.
This interest is connected to the high precision experiment at LEAR
(CERN) measuring the difference in the gravitational acceleration of
the proton and the antiproton \cite{PS-200}. For a review of earlier ideas
about antigravity
the reader is referred to the extensive article by Nieto and
Goldman \cite{GoldmanNieto} and the references therein.
The supergravity multiplet in the $N=2,8$ cases contains, in addition to
the graviton ($J=2$), a vector field $A_{\mu}^l$ ($J=1$). There
are also two Majorana gravitini ($J=\frac{3}{2}$) for $N=2$ \cite{Zachos}
and a scalar field $\sigma$ for $N=8$ \cite{Scherk,ScherkProc}.
The former fields are immaterial for our purposes and will be ignored in
the following. The field $\sigma$ does not induce any violation of the
equivalence principle. We comment below about the vanishing of the
scalar field contribution to
violations of the equivalence principle. It is also to be noted that there are
important differences between extended supergravity and the Standard Model, and
therefore the particles mentioned in this work should not be intended as the
objects familiar from the Standard Model.
The E\"otv\"os experiment forces upon us the assumption that the field
$A_{\mu}^l$ have a nonvanishing mass, which may have a dynamical origin
\cite{Scherk,ScherkProc}. In any case, the vector receives a mass
through the Higgs mechanism\setcounter{equation}{0}
\begin{equation} \label{1}
m_l=\frac{1}{R_l}=k\, m_{\phi}\langle \phi \rangle \; ,
\end{equation}
where the mass of the Higg
--like
field equals its (nonvanishing) vacuum expectation value ({\em v.e.v.})
\begin{equation} \label{2}
m_{\phi}=\langle \phi \rangle \; .
\end{equation}
Thus, Scherk's model of antigravity leads to the possibility of
violating the equivalence principle on a range of distances of order
$R_l$, where $R_l$ is the $A_{\mu}^l$ Compton wavelength. The available
limits set by the experimental tests of the equivalence principle
allow us to constrain the {\em v.e.v.} of the scalar field $\phi$, and
therefore its mass. It must be noted that the possibility of a
massless field $A_{\mu}^l$ was already ruled out by Scherk using the
E\"{o}tv\"{o}s experiments available at that time \cite{Scherk}.
In the present paper we review the limits provided by the present day
experiments, and those obtainable from experiments currently under
planning for
the near future. The Compton wavelength of
the gravivector thus obtained is of order 10~m, or less \cite{belfar}.
Therefore,
the concept of antigravity in the context of extended supergravity
cannot play any role in astrophysics, except possibly for processes
involving the strong gravity regime, i.e. near black holes or in the
early universe.
It is worth to remind the reader that there are interesting
connections between antigravity in $N=2,$~$8$ supergravities and {\em
CP} violation experiments, via the consideration of the
$K^0$--$\overline{K}^0$ system in the terrestrial gravitational field
\cite{Scherk}. However, the present experiments on {\em CP} violations
yield bounds on the range of the gravivector field which are less
stringent than those obtained from the tests of the equivalence
principle \cite{belfar}.
The present paper has the following structure: in Sec.~2 we recall the
basic features of Scherk's antigravity. The limits on the
Compton wavelength of the gravivector and the mass of the scalar
field deriving from the current experimental verifications of the
equivalence principle are discussed in Sec.~3 . Improvements coming from
experiments under planning are also considered. The conclusions are
presented in Sec.~4.
\section{Antigravity effects in $N>1$ supergravity}
In $N=2$,~$8$ supergravity theories, the gravivector field $A_{\mu}^l$
couples to the fields of the matter scalar multiplet with strengths
\begin{equation} \label{3}
g_i=\pm k\, m_i \end{equation}
\cite{Zachos} for $N=2$ and
\begin{equation}
g_i=\pm 2k\,m_i \end{equation}
\cite{SS,CremmerSS} for $N=8$. Here $k=(4\pi G)^{1/2}$, $m_i$ are the
quark and lepton masses, the positive and negative signs hold for
particles and antiparticles, respectively, and $g=0$ for
self--conjugated particles. As a consequence, in the interaction of an
atom with the gravitational field, the vector field $A_{\mu}^l$
``sees'' only the particles constituting the nucleon which are not
self--conjugated, while the graviton couples to the real mass of the
nucleon. The potential for an atom of atomic and mass numbers
($Z$,~$A$) in the static field of the Earth is \cite{Scherk}
\begin{equation} \label{4}
V=-\,\frac{G}{r} \left[ MM_{\oplus}-\eta M^0 {M^0_{\oplus}} \,
f\left( \frac{R_\oplus}{R_l} \right) \exp(-r/R_l) \right] \; , \end{equation}
where
\begin{equation} \label{5}
\eta =\left\{ \begin{array}{cllll}
1 & \,\,\,\,\,\, , & \;\;\; N=2 & & \nonumber \\
4 & \,\,\,\,\,\, , & \;\;\; N=8 & & \nonumber
\end{array} \right. \; .
\end{equation}
$R_l$ is the Compton wavelength of the gravivector, and
$R_\oplus =6.38 \cdot 10^6$~m, $ M_{\oplus}=5.98 \cdot 10^{24}$~kg
are the earth radius and mass, respectively. The presence of the
function
\begin{equation} \label{6}
f(x)=3 \,\, \frac{x\cosh x-\sinh x}{x^3} \end{equation}
expresses the fact that a spherical mass distribution cannot be
described by a point mass located at the
center of the sphere, as in the case of a coulombic potential. The
masses in (\ref{3}) are given by
\begin{eqnarray}
&& M=Z(M_p+m_e)+(A-Z)M_n \; , \label{7} \\
&& M^0=Z(2m_u+m_d+m_e)+(A-Z)(m_u+2m_d) \; , \label{8}
\end{eqnarray}
where $M_p$, $M_n$ and $m_e$ are the proton, neutron and electron masses,
respectively. We describe the Earth by means of the average atomic composition
$(Z_{\oplus},2Z_{\oplus})$ which gives, from (\ref{7}),~(\ref{8})
\begin{equation} \label{9}
M^0_{\oplus} \simeq \frac{3m_u+3m_d+m_e}{M_p+M_n} \, M_{\oplus} \; .
\end{equation}
In $N=2,8$ supergravities, one of the scalar fields has a nonzero
{\em v.e.v.} and, as a consequence, the vector field $A_{\mu}^l$ acquires a
mass, as described by (\ref{2}) (the impossibility of a massless
$A_{\mu}^l$ being proved in ref.~\cite{Scherk}). This leads
to a violation of the equivalence principle, expressed by the
difference between the accelerations of two atoms with numbers
$(Z,A)$ and $(Z',A')$ in the field of the Earth
\begin{equation} \label{10}
\frac{\delta \gamma}{\gamma}=\eta \, \frac{(3m_u+3m_d+m_e)(m_e+m_u-m_d)}
{M_n \, (M_p+M_n) } \left( \frac{Z'}{A'}-\frac{Z}{A} \right) f\left(
\frac{R_{\oplus}}{R_l}\right) \left( 1+\frac{R_{\oplus}}{R_l} \right)
\exp(-R_{\oplus}/R_l) \; . \end{equation}
In the spontaneously broken $N=8$ supergravity a graviscalar appears
together with the gravivector and gives a contribution to the
gravitational acceleration $\gamma $. However, this contribution has the same
sign for both particles and antiparticles, and thus cancels in the difference
$\delta \gamma $ measured in experiments on the equivalence principle. Hence,
the graviscalar does not contribute to violations of the equivalence
principle.
\section{Experimental constraints on antigravity}
In the E\"{o}tv\"{o}s--like experiment performed at the University of
Washington \cite{EotWash} (hereafter ``E\"{o}t--Wash'') the equivalence
principle was tested using berillium and copper and aluminum and
copper. The equivalence principle was verified with a precision
\begin{equation} \label{11}
\left| \frac{\delta \gamma}{\gamma} \right| < 10^{-11} \; .
\end{equation}
In ref.~\cite{belfar}, it was found that the Compton wavelength $R_l$ of the
gravivector satisfies the constraint
\begin{equation} R_l \leq 34 \, \eta^{-1} \;\; {\mbox m} \; .
\end{equation}
This justifies the use of the results of the E\"{o}t--Wash experiment, which
holds its validity for distances of the order $10^4$~m or less. Eq.~(\ref{11})
was used in ref.~\cite{belfar} to set a lower limit on the mass
of the Higgs--like particle:
\begin{equation} \label{12}
m_{\phi}>5 \, \eta^{1/2} \;\;\;\;\;\mbox{GeV} \; . \end{equation}
The Moscow experiment \cite{Moscow}, in spite of its higher precision,
provides a less stringent limit on $m_{\phi}$, due to the fact that it
verified the equivalence principle in the gravitational field of the
Sun, and (\ref{3}) has to be modified accordingly \cite{belfar}.
Here we make use also of the experiments aimed to detect deviations
from Newton's inverse square law, which can be seen as precise tests of the
equivalence principle. In these experiments it is customary to
parametrize the deviations from the Newtonian form with a Yukawa--like
correction to the Newtonian potential
\begin{equation} \label{13}
V(r)=-\,\frac{GM}{r} \left( 1+\alpha \, \mbox{e}^{-r/R_l} \right) \; .
\end{equation}
In the following, we assume that, in the context
of antigravity, the parameter $\alpha$ is given by the value computed
for the E\"{o}t--Wash
experiment performed using copper ($Z=29 $, $A=63.5 $) and berillium
($Z'=4 $, $A'=9.0 $), i.e.
\begin{equation} \label{14}
\alpha =\left\{ \begin{array}{cllll}
6.36 \cdot 10^{-4} & \,\,\,\,\,\, & \;\;\; (N=2) & & \nonumber \\
2.54 \cdot 10^{-3} & \,\,\,\,\,\, & \;\;\; (N=8) \;. & & \nonumber
\end{array} \right.
\end{equation}
For the materials that are likely to be used in these experiments, the
values of $\alpha$ differ from those of (\ref{14}) only for a
factor of order unity. Moreover, our final limits on $m_{\phi}$ depend
on the square root of $\alpha$. For these reasons, it is safe to use
the values (\ref{14}) of $\alpha$ in the following computations.
Equations (\ref{1}) and (\ref{2}) provide us with the relation
\begin{equation}
\frac{m_{\phi}( \mbox{new})}{{m_{\phi}}^*}=\left(
\frac{{R_l}^*}{R_l( \mbox{new})}\right)^{1/2} \; ,
\end{equation}
where ${m_{\phi}}^*=5 \eta^{1/2}$~GeV and ${R_l}^*=34 \eta^{-1}$~m are,
respectively, the
lower limit on the scalar field mass and the
upper limit on the Compton wavelength
of the vector $A_{\mu}^l$ derived in ref.~\cite{belfar}, and
$m_{\phi}( \mbox{new})$, $R_l( \mbox{new})$ are the new limits on the
same quantities coming from the references considered in the following.
The 2$\sigma$ limits of ref.~\cite{Spero} (see their fig.~3) allow the
range of values of $R_l$:
\begin{equation} \label{15}
R_l \leq 1 \: \mbox{cm} \:\:\:\: , \:\:\:\: R_l\geq 5 \: \mbox{cm}
\end{equation}
for $N=2$ and
\begin{equation} \label{16}
R_l \leq 0.5 \: \mbox{cm} \:\:\:\: , \:\:\:\: R_l\geq 16 \: \mbox{cm}
\end{equation}
for $N=8$. This corresponds to the allowed range for the mass of the
Higgs--like scalar field:
\begin{equation} \label{17}
m_{\phi} \leq 130 \: \mbox{GeV} \:\:\:\: , \:\:\:\: m_{\phi}\geq 292
\: \mbox{GeV} \:\:\:\:\:\:\:\: (N=2)
\end{equation}
\begin{equation} \label{18}
m_{\phi} \leq 73 \: \mbox{GeV} \:\:\:\: , \:\:\:\:\: m_{\phi}\geq 412 \:
\mbox{GeV} \:\:\:\:\:\:\:\: (N=8) \; .
\end{equation}
The curve~A of fig.~13 in ref.~\cite{Hoskinsetal} gives
\begin{equation}
R_l \leq 0.6 \: \mbox{cm} \:\:\:\: , \:\:\:\: R_l\geq 10 \: \mbox{cm}
\end{equation}
for $N=2$ and
\begin{equation}
R_l \leq 0.4 \: \mbox{cm} \:\:\:\: , \:\:\:\: R_l\geq 32 \: \mbox{cm}
\end{equation}
for $N=8$. Equivalently,
\begin{equation}
m_{\phi} \leq 92 \: \mbox{GeV} \:\:\:\: , \:\:\:\: m_{\phi}\geq 376
\: \mbox{GeV} \:\:\:\:\:\:\:\: (N=2)
\end{equation}
\begin{equation}
m_{\phi} \leq 52 \: \mbox{GeV} \:\:\:\: , \:\:\:\: m_{\phi}\geq 461 \:
\mbox{GeV} \:\:\:\:\:\:\:\: (N=8) \; .
\end{equation}
The null result of the Shternberg \cite{Shternberg} experiment reviewed by
Milyukov \cite{Milyukov} in the light of Scherk's work provides us
with the limits:
\begin{equation}
R_l \leq 4 \: \mbox{cm} \:\:\:\: , \:\:\:\: R_l\geq 13 \: \mbox{cm}
\end{equation}
for $N=2$ and
\begin{equation}
R_l \leq 2.2 \: \mbox{cm} \:\:\:\: , \:\:\:\: R_l\geq 40 \: \mbox{cm}
\end{equation}
for $N=8$. These are equivalent to:
\begin{equation}
m_{\phi} \leq 82 \: \mbox{GeV} \:\:\:\: , \:\:\:\: m_{\phi}\geq 146 \:
\mbox{GeV} \:\:\:\:\:\:\:\: (N=2)
\end{equation}
\begin{equation}
m_{\phi} \leq 46 \: \mbox{GeV} \:\:\:\: , \:\:\:\: m_{\phi}\geq 197 \:
\mbox{GeV} \:\:\:\:\:\:\:\: (N=8) \; .
\end{equation}
Therefore, the best available limits on the mass of the scalar field
are given by
\begin{equation} \label{19}
m_{\phi} \leq 82 \: \mbox{GeV} \:\:\:\: , \:\:\:\: m_{\phi}\geq
376 \: \mbox{GeV} \:\:\:\:\:\:\:\: (N=2)
\end{equation}
\begin{equation} \label{20}
m_{\phi} \leq 46 \: \mbox{GeV} \:\:\:\: , \:\:\:\: m_{\phi}\geq 461 \:
\mbox{GeV} \:\:\:\:\:\:\:\: (N=8) \; .
\end{equation}
A high precision test of the equivalence principle in the
field of the Earth is currently under planning in Moscow \cite{Kalebin}. The
precision expected to be achieved in this experiment is
\begin{equation} \label{21}
\left| \frac{\delta \gamma}{\gamma} \right| < 10^{-15} \; .
\end{equation}
Adopting the upper bound $R_l \leq 34 \, \eta^{-1}$~m found in
ref.~\cite{belfar} and approximating the function $f(x)$ for
$x=R_{\oplus}/R_l >>1$ as
\begin{equation} \label{22}
f(x) \simeq \frac{3}{2x^2} \, \mbox{e}^x \; ,
\end{equation}
we obtain
\begin{equation} \label{23}
\frac{ \left| \delta \gamma / \gamma \right|_{\mbox{future}}}
{\left| \delta \gamma / \gamma \right|_{\mbox{E\"{o}t-Wash}}}
=10^{-4} \; .
\end{equation}
In the case that the new experiment verifies the equivalence principle
with the expected accuracy, the limits on $m_{\phi}$ would be pushed
to
\begin{equation} \label{24}
m_{\phi} \geq 500 \, \eta^{1/2} \:\:\: \mbox{GeV} \; .
\end{equation}
\section{Conclusions}
The tests of the equivalence principle and the null results on
deviations from Newton's inverse square law provide constraints on the
mass of the Higgs--like boson appearing in extended supergravity
theories. We have reviewed the limits obtainable from the available
experiments in the context of $N=2,8$ supergravity, and we have discussed
also the impact on the field of the future high precision experiment
being planned in Moscow.
There have been many papers on the effects of non--Newtonian gravity in
astrophysics, in particular those due to a fifth force like the one obtainable
from $N=2,8$ supergravity in the weak field limit (see references in
\cite{GoldmanNieto}). However, the upper bound of $34 \eta^{-1}$~m on the
Compton wavelength $R_l$ of
the gravivector field found in ref.~\cite{belfar} implies that
antigravity does not play any role in nonrelativistic astrophysics since
the length scales involved in stellar\footnote{The conclusion that stellar
structure is unaffected by antigravity might change if the non--Newtonian
force alters the equation of state of the matter composing the star
\cite{stellar}.}, galactic and supergalactic
structures dominated by gravity are much larger than $R_l$. Antigravity
could affect, in principle, processes that take place in the
strong gravity regime, where smaller distance scales are involved.
Examples of these situations are processes occurring near black hole
horizons or in the early universe, when the size of the universe is
smaller than, or of the order of, $R_l$. The relevance of antigravity
in such situations will be studied in future publications.
Our final remark concerns a point that apparently went unnoticed in
the literature on supergravity: the detection of gravitational waves
expected in a not too far future will shed light on the
correctness of supergravity theories. In fact, after the dimensional
reduction is performed, the action of the theory contains scalar and
vector fields as well as the usual metric tensor associated to spin~2
gravitons \cite{ScherkProc}. These fields are responsible for the presence of
polarization modes in gravitational waves, the effect of which
differs from that of the spin~2 modes familiar from general
relativity. Therefore, extended supergravities and general
relativity occupy different classes in the $E(2)$ classification of
Eardley {\em et al.} \cite{Eardleyetal} of gravity theories. The extra
polarization states are detectable, in principle, in a gravitational
wave experiment employing a suitable array of detectors \cite{Eardleyetal}.
However, it must be noted that a detailed study of gravitational wave
generation taking into account the antigravity phenomenon is not available at
present. Such a work would undoubtedly have to face the remarkable
difficulties well known from the studies of gravitational wave generation in
the context of
general relativity.
\section*{Acknowledgments}
We are grateful to Prof. G. A. Lobov for drawing our attention to the ITEP
experiment. V.~F. acknowledges the warm hospitality of the INFN group in
Frascati, where this research was carried out.
{\small |
\section{Introduction}
\label{secintro}
The suggestion of Goldman, Henderson, and Thomas~\cite{ght92} that the
contribution of $\rho$-$\omega$ mixing to charge-symmetry-breaking
(CSB) observables is suppressed in the low momentum transfer regime
has opened the search for new sources of isospin violation. Since
then, many calculations, using a variety of models, have confirmed the
suppression of the $\rho$-$\omega$ mixing amplitude at small spacelike
momenta~\cite{piewil93,hats93,krein93,mitch94,oconn94,maltman95}.
Indeed, it has been shown that the $\rho$-$\omega$ mixing amplitude is
zero at $q^2=0$ in all models with vector mesons coupled to conserved
currents~\cite{oconn94}. Yet, in
Refs.~\cite{piewil93,hats93,krein93,mitch94,oconn94,maltman95} no
alternate mechanisms to $\rho$-$\omega$ mixing are proposed.
The phenomenological impact of this gap must be emphasized:
the CSB potential from $\rho$-$\omega$ mixing --- with the mixing amplitude
fixed at the omega-meson point --- accounts for some 40\% of
the difference between the neutron and proton analyzing powers ($\Delta A$)
measured in elastic $\vec{n}-\vec{p}$ scattering at 183 MeV~\cite{knut90}.
Without this contribution the previous agreement between theory and
experiment would be upset~\cite{miller86,willia87,miller90,iqnisk94}. Although
the suppression of the mixing amplitude continues to be
controversial~\cite{miller94,oconn95}, sources of additional isospin
violation are interesting in their own right and deserve examination.
Indeed, the aim of the present paper is to show that a recently
proposed CSB mechanism --- based on isospin-violating meson-nucleon
coupling constants~\cite{ghp95} --- is sufficient to restore
the agreement with experiment. Specifically, we examine the effect
of these new sources of CSB on the spin-singlet--triplet mixing
angles; these are the fundamental dynamical quantities driving
$\Delta A$~\cite{willia87}.
Most theoretical efforts devoted to understanding CSB
observables use a nucleon-nucleon ($NN$) interaction
constrained by two-nucleon
data~\cite{miller86,willia87,holz87}. In such a picture, isospin
violations arise from electromagnetic effects and hadronic mass
differences. Sources of CSB can be classified in terms of three
distinct contributions: (i) isovector-isoscalar mixing in the
meson propagator, (ii) isospin-breaking in the nucleon wave
function, and (iii) isospin breaking in the meson-nucleon and
photon-nucleon vertices. Rho-omega mixing, the
proton-neutron mass difference, and the difference between
the electric charge of the proton and the neutron are typical
examples of (i), (ii), and (iii), respectively.
The existence of isovector-isoscalar mixing, such as $\pi$-$\eta$ and
$\rho$-$\omega$ mixing, is well established. For example,
$\rho$-$\omega$ mixing has been observed experimentally in
$e^{+}e^{-}\rightarrow\pi^{+}\pi^{-}$ measurements at the
$\omega-$meson production point~\cite{barkov85}. However,
the suggested suppression of the mixing amplitudes at small
spacelike momenta lessens their impact on
CSB observables. This is, in part, why we consider other
sources of isospin violation in this paper.
Isospin breaking in the nucleon wave function in a hadronic model
is driven by the neutron-proton mass difference. Indeed, it is through
this mechanism that charged-pion exchange dominates~\cite{cheung80,gersten81}
the class IV potential~\cite{henmil79} at moderate momentum transfers.
Isospin breaking in the nucleon wave function can also arise in a quark
model picture from the mixing of the nucleon to
$|J^{\pi}=1/2^{+};T=3/2\rangle$ baryon states~\cite{dmitra95}. While
undoubtedly nonzero, one expects the $T=3/2$ components of the nucleon
to be small due to the large mass difference between the nucleon and
the $\Delta(1910)$ --- the first $P_{31}$ baryon. In contrast, the
$\rho$-$\omega$ mass difference is a mere 12 MeV. Thus, we turn to
the meson-nucleon coupling constants as the possible sources of
isospin violation demanded by data.
While there have been calculations of isospin-violating meson-nucleon
coupling constants~\cite{miller90,mitra67,thomas81,henzha87,iqnisk88}, their
impact on class IV CSB observables has only recently been
considered~\cite{ghp95}. Here, as earlier~\cite{ghp95}, we adopt a
nonrelativistic quark model to calculate isospin breaking in the
meson-nucleon coupling constants. In the model the coupling constants
emerge from evaluating matrix elements of quark currents of the
appropriate Lorentz and flavor structure between nucleon states. The
isospin violations arise from the up-down quark mass difference.
Radiative corrections to the vertices have also been evaluated
and are found to be small~\cite{morris68,yakin79}. Here we study
the phenomenological impact of Ref.~\cite{ghp95} in greater detail.
In order to do this, we estimate the $q^2$ dependence of the
isospin-breaking found in the vertices at $q^2=0$.
We have organized the paper as follows. In Sec.~\ref{secivcc}
the model is introduced, and isospin-violating meson-nucleon
coupling constants are computed. We show that in the $q^2=0$ limit
the couplings depend merely on the spin-flavor structure of the
nucleon wave function; they are insensitive to the spatial components
of the nucleon wave function. In Sec.~\ref{seccsbp} we use these
findings to compute the resulting CSB potentials. In particular, we
obtain a large contribution from omega-meson exchange to the class IV
potential. We quantify the impact of isospin-violation in the
$NN\omega$ vertex by computing the resulting spin-singlet--triplet
mixing angles --- these are the basic building blocks of $\Delta A$.
These results are presented in Sec.~\ref{secresults}. Finally, we
discuss the impact of our work in Sec.~\ref{secconcl}.
\section{Isospin-violating meson-nucleon coupling constants}
\label{secivcc}
We are interested in computing the coupling of an on-shell nucleon
to the neutral mesons $\omega$, $\rho^{0}$, $\pi^{0}$, and $\sigma$.
The off-shell vertices could engender additional isospin breaking,
but our primary focus is on the $NN$ system, so that we will not
consider these effects further. The exchanged mesons couple to nucleon
currents of the appropriate Lorentz character, and the meson-nucleon
coupling constants emerge from evaluating the matrix elements of these
currents in the quark model. The difference in the up and down constituent
quark masses thus gives rise to isospin-violating meson-nucleon coupling
constants. At $q^{2}=0$ these couplings are determined from the spin and
flavor structure of the nucleon wave function alone. In contrast, the
couplings at $q^2=0$ of the nucleon to the charged mesons are sensitive
to the quark momentum distribution as well, and are, therefore, more
model dependent~\cite{miller90}. We shall consider the
neutral-vector-meson--nucleon vertices first, as they are relevant to
the $\Delta A$ measurement. The most general form for these on-shell
$NN$-meson vertex functions, consistent with Lorentz covariance
and parity invariance, are
\begin{mathletters}
\label{vert}
\begin{eqnarray}
-ig_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}
\Lambda^{\mu}_{\lower 2pt \hbox{$\scriptstyle NN\omega$}} &=&
-ig_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}
\left(
g^{\omega}_N \gamma^{\mu} +
if^{\omega}_N \sigma^{\mu\nu} {(p'-p)_{\nu} \over 2M_{N}}
\right) \;, \\
-ig_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
\Lambda^{\mu}_{\lower 2pt \hbox{$\scriptstyle NN\rho$}} &=&
-ig_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
\left(
g^{\rho}_N \gamma^{\mu} +
if^{\rho}_N \sigma^{\mu\nu} {(p'-p)_{\nu} \over 2M_{N}}
\right) \;,
\end{eqnarray}
\end{mathletters}
where $g_{\lower 2pt \hbox{$\scriptstyle NN\alpha$}}$
($\alpha=\omega,\rho$) are the isospin-averaged, phenomenological,
meson-nucleon coupling constants, determined from
fits to the $NN$ phase shifts and to the properties of the
deuteron~\cite{machl87,machl89}, and $M_{N}$ is the nucleon mass.
We compute the couplings, that is, $g^{\alpha}_{N}(q^2)$ and
$f^{\alpha}_{N}(q^2)$, by assuming that the $NN\alpha$ vertex functions
can be related to the matrix elements of quark currents of the
appropriate Lorentz and flavor structure between nucleon states,
computed in the nonrelativistic quark model. Thus,
\begin{mathletters}
\begin{eqnarray}
\langle N(p',s') | J^{\mu;+} | N(p,s) \rangle &=&
\bar{U}(p',s')
\Lambda^{\mu}_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}
U(p,s) \;, \\
\langle N(p',s') | J^{\mu;-} | N(p,s) \rangle &=&
\bar{U}(p',s')
\Lambda^{\mu}_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
U(p,s) \;.
\end{eqnarray}
\label{vertex}
\end{mathletters}
Here $U(p,s)$ denotes a on-shell nucleon spinor of mass $M_{N}$,
momentum $p$, and spin $s$. We shall focus on the couplings at
$q^2=0$, where $q\equiv p'-p$, as the nonrelativistic quark model
is best suited to an estimate in the static limit. The quark currents
$J^{\mu;\pm}$ are
\begin{mathletters}
\begin{eqnarray}
J^{\mu;+} &=& {1\over 3}\, \bar{u} \gamma^{\mu} u
+ {1\over 3}\, \bar{d} \gamma^{\mu} d \;, \\
J^{\mu;-} &=& \bar{u} \gamma^{\mu} u
- \bar{d} \gamma^{\mu} d \;.
\end{eqnarray}
\label{veccurrent}
\end{mathletters}
It is the quark vector current which is appropriate to the
vector-meson--nucleon vertex; the second superscript ($\pm$)
denotes its symmetry under the $u \leftrightarrow d$ flavor
transformation. Note that the constituent quarks are assumed
to be elementary: no quark form factors have been introduced.
The isoscalar vector quark charge is $1/3$, whereas
the isovector vector quark charge is $+1$ for the up quark and
$-1$ for the down quark. The charge assignments are made such
that $g_N^{\omega}=1$, $g_p^{\rho}=1$, and $g_n^{\rho}=-1$,
at $q^2=0$.
Our model stems from the notion of vector dominance~\cite{nambu57}.
Vector dominance presumes that a photon's interaction with a nucleon
is mediated by the rho --- or omega --- meson. Here we argue that
the coupling of the vector mesons themselves to the nucleon can be
determined via matrix elements of the appropriate isospin
components of the quark {\it vector} current.
Our model does not predict the isospin-conserving coupling
constants $g_{\lower 2pt \hbox{$\scriptstyle NN\alpha$}}$; these must
be extracted from phenomenological fits to two-nucleon data. However,
the isospin-violating pieces, as well as the
tensor-to-vector ratio, can be calculated within the model.
Note that the vector dominance nature of our model implies that the
quarks couple to {\it conserved currents}. We estimate the resulting
coupling constants using the nonrelativistic quark model (NRQM); this
is an additional assumption.
The couplings $g^{\alpha}_{N}$ and $f^{\alpha}_{N}$ are
functions of the meson four-momentum $q^{2}$,
though we shall focus on the couplings at
$q^2=0$. In this limit the couplings are
insensitive to the spatial component of the nucleon wave function; they
follow directly from its spin and flavor content alone.
In the $SU(6)$ limit~\cite{perkins87},
\begin{eqnarray}
|p \!\uparrow \rangle = {1 \over \sqrt{18}}
\Big(
&& 2|u\!\uparrow u\!\uparrow d\!\downarrow\rangle -
|u\!\uparrow u\!\downarrow d\!\uparrow\rangle -
|u\!\downarrow u\!\uparrow d\!\uparrow\rangle + \nonumber \\
&& 2|u\!\uparrow d\!\downarrow u\!\uparrow\rangle -
|u\!\downarrow d\!\uparrow u\!\uparrow\rangle -
|u\!\uparrow d\!\uparrow u\!\downarrow\rangle + \\
&& 2|d\!\downarrow u\!\uparrow u\!\uparrow\rangle -
|d\!\uparrow u\!\uparrow u\!\downarrow\rangle -
|d\!\uparrow u\!\downarrow u\!\uparrow\rangle \nonumber
\Big) \;.
\end{eqnarray}
The neutron spin-up wave function, $|n\!\uparrow \rangle$, is obtained
by exchanging the up and down quarks in the expression for
$|p \!\uparrow \rangle$. The isospin violations arise from the
difference in the up and down constituent quark masses. The couplings
constants are obtained from computing the matrix elements found in the
nonrelativistic reduction of Eq.~(\ref{vertex}) in the quark model;
{\it i.e.},
\begin{mathletters}
\label{themodel}
\begin{eqnarray}
g^{\omega}_N = \sum_{i=1}^3
g^{+}_i
\langle N \!\uparrow | 1 | N\! \uparrow\rangle \quad &;& \quad
{(g^{\omega}_N + f^{\omega}_N) \over {2M_N}} = \sum_{i=1}^3
\mu^{+}_i
\langle N \!\uparrow | \sigma^z_i | N\! \uparrow\rangle \;, \\
g^{\rho}_N = \sum_{i=1}^3
g^{-}_i
\langle N \!\uparrow | 1 | N\! \uparrow\rangle \quad &;& \quad
{(g^{\rho}_N + f^{\rho}_N) \over {2M_N}} = \sum_{i=1}^3
\mu^{-}_i
\langle N \!\uparrow | \sigma^z_i | N\! \uparrow\rangle \;.
\end{eqnarray}
\end{mathletters}
Note that we have introduced the quark magnetic moment
$\mu^{\pm}_{i} \equiv g^{\pm}_{i}/2m_{i}$, with the
charges $g^{\pm}_{i}$ given in Table~\ref{tableone}.
In the following presentation we discuss only the coupling
of the nucleon to the $\omega-$meson, as an illustrative
example. Our results, collected in Table~\ref{tabletwo},
include the couplings to the other mesons as well.
The vector coupling of the $\omega-$meson to the nucleon is
determined by simply counting the quark charges:
\begin{mathletters}
\begin{eqnarray}
g^{\omega}_p &=&
2g^{+}_u + g^{+}_d = 1 \;, \\
g^{\omega}_n &=&
2g^{+}_d + g^{+}_u = 1 \;.
\end{eqnarray}
\label{model1}
\end{mathletters}
The tensor coupling, in contrast, depends on the spin structure
of the nucleon wave function:
\begin{mathletters}
\begin{eqnarray}
\mu^{\omega}_p \equiv
{g^{\omega}_p + f^{\omega}_p \over {2M_p}} &=&
{4 \over 3} \mu^{+}_u - {1 \over 3} \mu^{+}_d =
{1 \over 18} \left(
{4 \over m_{u}} - {1 \over m_{d}} \right) \;, \\
\mu^{\omega}_n \equiv
{g^{\omega}_n + f^{\omega}_n \over {2M_n}} &=&
{4 \over 3} \mu^{+}_d - {1 \over 3} \mu^{+}_u =
{1 \over 18} \left(
{4 \over m_{d}} - {1 \over m_{u}} \right) \;.
\end{eqnarray}
\label{model2}
\end{mathletters}
It is
useful to construct
isoscalar and isovector combinations at the {\it nucleon} level; i.e.,
\begin{eqnarray}
g^{\omega}_N &=&
g^{\omega}_p {1 \over 2}(1+\tau_z) +
g^{\omega}_n {1 \over 2}(1-\tau_z) \equiv
g^{\omega}_0 + g^{\omega}_1 \tau_z \;, \\
\mu^{\omega}_N &=&
\mu^{\omega}_p {1 \over 2}(1+\tau_z) +
\mu^{\omega}_n {1 \over 2}(1-\tau_z) \equiv
\mu^{\omega}_0 + \mu^{\omega}_1 \tau_z \;,
\end{eqnarray}
where
\begin{eqnarray}
\label{omegag}
g^{\omega}_0 + g^{\omega}_1 \tau_z &=& 1 \;, \\
\mu^{\omega}_0 + \mu^{\omega}_1 \tau_z &=&
\label{omegaf}
{1 \over 6m} \left[ 1 + {5\over 6}{\Delta m \over m} \tau_z \right] \equiv
\left[ {(g^{\omega}_0 + f^{\omega}_0) \over 2M} +
{(g^{\omega}_1 + f^{\omega}_1) \over 2M} \tau_z \right]
\end{eqnarray}
with
\begin{equation}
M \equiv {1 \over 2}(M_{n}+M_{p}) \;; \quad
m \equiv {1 \over 2}(m_{d}+m_{u}) \;; \quad
\Delta m \equiv (m_{d}-m_{u}) \;.
\end{equation}
The expression in Eq.~(\ref{omegaf}) is given to leading order in
$\Delta m /m$ only. Note that isospin breaking in the $f$ and $g$
couplings is realized in the $f$ {\it alone} and that the breaking
in the $\omega$ tensor coupling is isovector in character. The
$\omega$ --- and $\rho$ --- vector couplings are isospin-conserving.
The isospin-breaking in our model is connected to that of the
electromagnetic form factors through our assumption of vector
dominance; charge conservation protects the charge form factor
from isospin-breaking at zero momentum transfer~\cite{dmitra95}.
The tensor coupling is explicitly sensitive to the quark mass,
as seen in Eq.~(\ref{themodel}), and the isospin-breaking
corrections are generated by the up-down mass difference.
In the constituent quark model $\Delta m > 0$~\cite{licht89};
the up quark, which is lighter, has a larger magnetic moment than
the down quark. Henceforth we shall adopt the choice $M/3m \equiv 1$
in reporting the coupling constants. Our results are summarized
in Table~\ref{tabletwo}.
We now consider the isospin-conserving results. We find for the
tensor-to-vector ratio that
\begin{equation}
{f^{\omega}_{0} \over g^{\omega}_N} =0 \;; \quad
{f^{\rho}_{1} \over g^{\rho}_N} =4 \;.
\end{equation}
These results are qualitatively consistent with the $f^V_N/g^V_N$
ratios which emerge from phenomenological fits to the $NN$
interaction~\cite{machl87,machl89} --- recall that the Bonn B
potential parameters~\cite{machl89}, for example, are
$f^{\omega}_N/g^{\omega}_N=0$ and $f^{\rho}_N/g^{\rho}_N=6.1$.
This consistency is intimately connected to the
NRQM's ability to describe the nucleon magnetic
moments and to our assumption of vector dominance. In the NRQM,
with $M=3m$, the anomalous magnetic moment is purely isovector:
$\kappa_p=2$ and $\kappa_n=-2$. Note that $\kappa_p^{\rm exp}=1.79$
and $\kappa_n^{\rm exp}=-1.91$. These successes gives us confidence
in using our model to compute the isospin-violating corrections to
these coupling constants.
For completeness, we shall now consider isospin breaking in the
$NN\sigma$ and $NN\pi^0$ vertices as well. We exclude the $NN\eta$
vertex from this discussion because the
$g_{\lower 2pt \hbox{$\scriptstyle NN\eta$}}$ coupling constant
is poorly constrained by $NN$ data~\cite{coon82}. The appropriate
vertex functions are
\begin{mathletters}
\begin{eqnarray}
\label{vertsig}
ig_{\lower 2pt \hbox{$\scriptstyle NN\sigma$}}
\Lambda^{\rm s}_{\lower 2pt \hbox{$\scriptstyle NN\sigma$}} &=&
ig_{\lower 2pt \hbox{$\scriptstyle NN\sigma$}}
\left(
g^{\sigma}_N
\right) {1} \;, \\
g_{\lower 2pt \hbox{$\scriptstyle NN\pi$}}
\Lambda^{5}_{\lower 2pt \hbox{$\scriptstyle NN\pi$}} &=&
\phantom{i}
g_{\lower 2pt \hbox{$\scriptstyle NN\pi$}}
\left(
g^{\pi}_N
\right)\gamma^{5} \;.
\end{eqnarray}
\end{mathletters}
We have assumed pseudoscalar, rather than pseudovector, coupling
for the pion in order to be consistent with earlier calculations
of charge-symmetry
breaking~\cite{miller86,willia87,holz87,cheung80,gersten81}.
In our model, we connect the vertex functions
to matrix elements of quark currents, so that
\begin{mathletters}
\begin{eqnarray}
\langle N(p',s') | J^{\rm s} | N(p,s) \rangle &=&
\bar{U}(p',s')
\Lambda^{\rm s}_{\lower 2pt \hbox{$\scriptstyle NN\sigma$}}
U(p,s) \;, \\
\langle N(p',s') | J^{5} | N(p,s) \rangle &=&
\bar{U}(p',s')
\Lambda^{5}_{\lower 2pt \hbox{$\scriptstyle NN\pi$}}
U(p,s) \;,
\end{eqnarray}
\label{vertexsig}
\end{mathletters}
where
\begin{mathletters}
\begin{eqnarray}
J^{\rm s}(q) &=& {1\over 3}\, \bar{u} u
+ {1\over 3}\, \bar{d} d \;, \\
J^{5}(q) &=& {1\over 5}\, \bar{u} \gamma^{5} u
- {1\over 5}\, \bar{d} \gamma^{5} d \;.
\end{eqnarray}
\label{scurrent}
\end{mathletters}
The charges $g^{\rm s}_i=1/3$, $g^5_u=1/5$, and $g^5_d=-1/5$
have been chosen such that $g_N^{\sigma}=1$, $g_p^{\pi}=1$, and
$g_n^{\pi}=-1$ when $\Delta m=0$. Evaluating the nonrelativistic
reduction of Eqs.~(\ref{vertexsig}a) and (\ref{vertexsig}b)
in the quark model, we find
\begin{mathletters}
\label{thesigpi}
\begin{eqnarray}
g^{\sigma}_N &=& \sum_{i=1}^3
g^{{\rm s}}_i
\langle N \!\uparrow | 1 | N\! \uparrow\rangle \;, \\
{g^{\pi}_N \over {2M_N}} &=& \sum_{i=1}^3
\mu^{5}_i
\langle N \!\uparrow | \sigma^z_i | N\! \uparrow\rangle \;,
\end{eqnarray}
\end{mathletters}
where we have defined $\mu^5_i\equiv g^5_i/2m_i$. From
Eq.~(\ref{thesigpi}a) we see that the sigma meson generates
merely a spin-independent coupling to the nucleon in the
nonrelativistic limit, so that there is no isospin breaking
in the $NN\sigma$ vertex and no contribution from sigma exchange
to the CSB potential. Thus, we will not consider sigma exchange
further. However, the quark mass dependence contained in $\mu^5_i$
in Eq.~(\ref{thesigpi}b) implies that the isospin-breaking in the
pion case is finite. The breaking to ${\cal O}(\Delta m/m)$ is
indicated in Table~\ref{tabletwo}. Note, however, that the computed
breaking at $q^2=0$ depends on the nature of the assumed pion-nucleon
coupling. If we had chosen {\it pseudovector} coupling, rather,
then no isospin breaking would result. The pseudovector current
contains no quark mass dependent pieces in the nonrelativistic limit.
Thus, our prediction in the $\pi^0$ case is decidedly more model
dependent than in the $\rho^0$ and $\omega$ channels. Moreover,
in the latter case, the compatibility of the computed tensor-to-vector
coupling constant ratios with the Bonn potential indicates that vector
dominance, which we assume, has some phenomenological support. Note that
in the $\pi^0$ case, there is no such independent support of our
``pseudoscalar dominance'' assumption. This concludes our discussion
of isospin breaking in the $NN$-meson vertices.
\section{Charge symmetry breaking potentials}
\label{seccsbp}
We shall now compute the CSB potentials which arise from the
isospin-violating couplings computed in the previous section
and tabulated in Table~\ref{tabletwo}. In an one-boson
exchange approximation, presuming the form of the
isospin breaking found in the $q^2=0$ results,
we obtain the following CSB potentials
for $\omega$, $\rho^0$, and $\pi^0$ exchange, respectively:
\begin{mathletters}
\begin{eqnarray}
\widehat{V}^{\omega}_{\rm CSB} &=& {V}^{\omega}_{\rm CSB}
\Big[ \Gamma^{\mu}(1) \gamma_{\mu}(2) \tau_z(1) +
\gamma^{\mu}(1) \Gamma_{\mu}(2) \tau_z(2) \Big]
\;, \label{vomegaa} \\
\widehat{V}^{\rho}_{\rm CSB} &=& {V'}^{\rho}_{\rm CSB}
\Gamma^{\mu}(1) \Gamma_{\mu}(2) \Big[ \tau_z(1) + \tau_z(2) \Big] +
{V}^{\rho}_{\rm CSB}
\Big[ \Gamma^{\mu}(1) \gamma_{\mu}(2) \tau_z(2) +
\gamma^{\mu}(1) \Gamma_{\mu}(2) \tau_z(1) \Big]
\;, \label{vrhoa} \\
\widehat{V}^{\pi}_{\rm CSB} &=& {V}^{\pi}_{\rm CSB}
\gamma^{5}(1) \gamma^{5}(2) \Big[ \tau_z(1) + \tau_z(2) \Big]
\;, \label{vpiona}
\end{eqnarray}
\end{mathletters}
where $\Gamma^{\mu} \equiv i\sigma^{\mu\nu}(p'-p)_{\nu}/2M$
and we have defined
\begin{mathletters}
\begin{eqnarray}
{V}^{\omega}_{\rm CSB}(q) &\equiv&
-\left({g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}^2
\over q^2 - m_{\omega}^2}\right)f^\omega_1 g^\omega_0
\;, \label{vomegab} \\
{V}^{\rho}_{\rm CSB}(q) &\equiv&
-\left({g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}^2
\over q^2 - m_{\rho}^2}\right)f^\rho_0 g^\rho_1
\;, \label{vrhob} \\
{V'}^{\rho}_{\rm CSB}(q) &\equiv&
-\left(
{g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}^2
\over q^2 - m_{\rho}^2 }\right)f^\rho_0 f^\rho_1
\;, \label{vprhob} \\
{V}^{\pi}_{\rm CSB}(q) &\equiv&
-\left({g_{\lower 2pt \hbox{$\scriptstyle NN\pi$}}^2
\over q^2 - m_{\pi}^2}\right)g^\pi_0 g^\pi_1
\;. \label{vpionb}
\end{eqnarray}
\end{mathletters}
Isospin breaking in the meson-nucleon vertices give rise to
the above CSB potentials, as per Eqs.~(\ref{vomegab})-(\ref{vpionb}).
The isospin-conserving tensor coupling is nonzero in the case of the
$\rho$ vertex, so that an additional potential of strength
${V'}_{\rm CSB}^{\rho}(q)$ arises. These contributions have been
considered only recently~\cite{ghp95}. Yet the potentials of
Eqs.~(\ref{vomegaa}) and (\ref{vpiona}) are identical in form to
those generated by $\rho$-$\omega$ and $\pi$-$\eta$ mixing, respectively.
That is,
\begin{mathletters}
\begin{eqnarray}
\widehat{V}^{\rho\omega}_{\rm CSB} &=& {V}^{\rho\omega}_{\rm CSB}
\Big[ \Gamma^{\mu}(1) \gamma_{\mu}(2) \tau_z(1) +
\gamma^{\mu}(1) \Gamma_{\mu}(2) \tau_z(2) \Big]
\;, \label{vrhoomegaa} \\
\widehat{V}^{\pi\eta}_{\rm CSB} &=& {V}^{\pi\eta}_{\rm CSB}
\gamma^{5}(1) \gamma^{5}(2) \Big[ \tau_z(1) + \tau_z(2) \Big]
\;, \label{vpietaa}
\end{eqnarray}
\end{mathletters}
where
\begin{mathletters}
\begin{eqnarray}
{V}^{\rho\omega}_{\rm CSB}(q) &\equiv& -
{f_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}
\over (q^2 - m_{\rho}^2 )(q^2 - m_{\omega}^2 )}
\langle \rho | H | \omega \rangle \;,
\label{vrhoomegab} \\
{V}^{\pi\eta}_{\rm CSB}(q) &\equiv& -
{g_{\lower 2pt \hbox{$\scriptstyle NN\pi$}}
g_{\lower 2pt \hbox{$\scriptstyle NN\eta$}}
\over (q^2 - m_{\pi}^2 )(q^2 - m_{\eta}^2 )}
\langle \pi | H | \eta \rangle \;.
\label{vpietab}
\end{eqnarray}
\end{mathletters}
Note that in Eq.~(\ref{vrhoomegab}) we introduce
$f_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}$, the
phenomenological tensor coupling of the Bonn model~\cite{machl89}.
Rather than performing a nonrelativistic reduction of the
potentials in Eqs.~(\ref{vomegaa})-(\ref{vpiona}) and
Eqs.~(\ref{vrhoomegaa})-(\ref{vpietaa}), we simply classify the
former as, either, ``$\rho\omega-\!\!$~like'' or ``$\pi\eta-\!\!$~like''
potentials. The effect of these new isospin-violating potentials on
CSB observables can then be readily elucidated. For example, the
contribution from omega-meson exchange is identical in structure to
that from $\rho$-$\omega$ mixing and thus contributes as well
to $\Delta A$ in elastic $\vec{n}-\vec{p}$ scattering. Indeed,
we now show that the contribution from omega-meson exchange is
comparable in magnitude and identical in sign to the one obtained
from $\rho$-$\omega$ mixing --- if the mixing amplitude is fixed at
its on-shell value.
\subsubsection{One-boson exchange potentials of the $\rho$-$\omega$ kind}
Potentials of the form given in Eq.~(\ref{vrhoomegaa}) give rise to
class III and class IV CSB potentials. They are generated by the
interference between the isospin-conserving vector coupling and the
isospin-violating tensor coupling; note, for example, Eq.~(\ref{vomegaa})
and the second term in Eq.~(\ref{vrhoa}). Unlike the case of the omega,
the isospin structure of rho exchange is not identical to that of
$\rho$-$\omega$ mixing; they are related by exchanging $\tau_{z}(1)
\leftrightarrow \tau_{z}(2)$. Thus, rho exchange contributes to the
class IV $\rho$-$\omega$ mixing potential with a sign opposite to that
of the omega. No sign changes are necessary when computing its
$\pi\eta-\!\!$~like or class III $\rho\omega-\!\!$~like contribution.
Note that the contribution from rho exchange is small relative to that
from the omega --- this emerges despite the larger isospin-violating
coupling associated with the rho vertex (see Table~\ref{tabletwo}).
The vector $NN\rho$ coupling is simply small relative to that of the
omega; in the Bonn potential
$g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}^2 /
g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}^2 \approx 27$~\cite{machl89}.
The relative importance of the various contributions can be estimated
by computing the CSB potentials at $q^2=0$. Recall that in this limit
the isospin-violating couplings are insensitive to the quark momentum
distribution; they depend only on the spin-flavor symmetry of the wave
function. Using the Bonn B potential parameters of Table~\ref{tablethree}
and a value for the quark-mass difference of
$\Delta m=4.1$~MeV~\cite{licht89}, we obtain the following results
at $q^2=0$:
\begin{mathletters}
\begin{eqnarray}
{V}^{\omega}_{\rm CSB}(q^2=0) &=&
{g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}^2
\over m_{\omega}^2} f^\omega_1 g^\omega_0
\approx 2.49~{\rm GeV}^{-2} \;, \\
{V}^{\rho}_{\rm CSB}(q^2=0) &=&
{g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}^2
\over m_{\rho}^2} f^\rho_0 g^\rho_1
\approx 0.18~{\rm GeV}^{-2} \;. \\
{V}^{\rho\omega}_{\rm CSB}(q^2=0) &=& -
{f_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}
\over m_{\rho}^2 m_{\omega}^2}
\langle \rho | H | \omega \rangle \Big|_{q^2=0}
=0 \;,
\end{eqnarray}
\end{mathletters}
Several remarks are in order. First, the $\rho$-$\omega$ mixing
amplitude, if modeled via fermion loops~\cite{piewil93,oconn94},
necessarily vanishes at $q^2=0$ in our model. Our model assumes
vector dominance, so that the vector-meson--nucleon vertices are
determined by the appropriate isospin components of the quark
electromagnetic current. Thus, the vector mesons couple to
currents that are conserved at the nucleon level, so that the
above result follows~\cite{piewil93,oconn94}. At the $q^2=0$
point, the charge-symmetry violation in our model comes
purely from the vertex contributions. Note that the rho meson
contribution to the latter is, indeed, small. It represents
merely a 7\% correction to the contribution from one-omega
exchange. Second, the strength of the CSB potentials generated
from omega exchange is comparable in magnitude to those obtained
from $\rho$-$\omega$ mixing if the {\it on-shell} value of the
mixing amplitude is assumed,
$\langle\rho|H|\omega\rangle|_{q^2=m_\omega^2}=-4520 \pm
600~{\rm MeV}^2$~\cite{coon87}.
Note, moreover, that the $\omega$ and on-shell $\rho$-$\omega$
mixing contributions are {\it identical} in sign. Specifically,
\begin{equation}
{\widetilde V}^{\rho\omega}_{\rm CSB}(q^2=0) = -
{f_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}
\over m_{\rho}^2 m_{\omega}^2}
\langle \rho | H | \omega \rangle \Big|_{q^2=m_\omega^2}
\approx 2.07~{\rm GeV}^{-2} \;,
\end{equation}
A CSB potential of this magnitude is needed for a successful
description of $\Delta A$ at 183 MeV~\cite{knut90}. Summing
our omega exchange contribution to the CSB potential and that
from on-shell $\rho$-$\omega$ mixing is not only internally
inconsistent but also gives a final potential which is too
large to fit the data --- see Sec.~\ref{secresults}. Our results
suggest that a class IV potential of the appropriate size is
generated by isospin-violations in the $NN\omega$ vertex, together
with small corrections from rho exchange and off-shell
$\rho$-$\omega$ mixing. This is the central result of our paper.
\subsubsection{One-boson exchange potentials of the $\pi$-$\eta$ kind}
Potentials of the form given in Eq.~(\ref{vpietaa}) generate class III
CSB potentials exclusively. The Lorentz structure of the first term in
Eq.~(\ref{vrhoa}) differs from that of the $\pi$-$\eta$ mixing and
one-pion exchange potentials, so that it is convenient to perform a
nonrelativistic reduction of all three contributions, i.e.,
\begin{mathletters}
\begin{eqnarray}
\widehat{V}^{\pi\eta}_{\rm CSB} &=& - {V}^{\pi\eta}_{\rm CSB}(q)
\left( {{\bf q}^2 \over 12M^2} \right)
\Big[ \mbox{\boldmath$\sigma$}_1 \cdot \mbox{\boldmath$\sigma$}_2 +
S_{12}(\hat{\bf q}) \Big]
\Big[ \tau_z(1) + \tau_z(2) \Big] \;, \label{vpietac} \\
\widehat{V}^{\pi}_{\rm CSB} &=& - {V}^{\pi}_{\rm CSB}(q)
\left( {{\bf q}^2 \over 12M^2} \right)
\Big[ \mbox{\boldmath$\sigma$}_1 \cdot \mbox{\boldmath$\sigma$}_2 +
S_{12}(\hat{\bf q}) \Big]
\Big[ \tau_z(1) + \tau_z(2) \Big] \;, \label{vpionc} \\
\widehat{V}^{\prime\rho}_{\rm CSB} &=& - {V}^{\prime\rho}_{\rm CSB}(q)
\left( {{\bf q}^2 \over 12M^2} \right)
\Big[ 2\mbox{\boldmath$\sigma$}_1 \cdot \mbox{\boldmath$\sigma$}_2 -
S_{12}(\hat{\bf q}) \Big]
\Big[ \tau_z(1) + \tau_z(2) \Big] \;, \label{vrhoc}
\end{eqnarray}
\end{mathletters}
where we have introduced the tensor operator $S_{12}(\hat{\bf q}) =
[3(\mbox{\boldmath$\sigma$}_1\cdot{\hat{\bf q}})
(\mbox{\boldmath$\sigma$}_2\cdot{\hat{\bf q}}) -
\mbox{\boldmath$\sigma$}_1\cdot\mbox{\boldmath$\sigma$}_2]$.
We estimate the relative size of these contributions by evaluating them
at $q^2=0$, noting table~\ref{tablethree}:
\begin{mathletters}
\label{zeroq2pieta}
\begin{eqnarray}
{V}^{\pi}_{\rm CSB}(q^2=0) &=&
{g_{\lower 2pt \hbox{$\scriptstyle NN\pi$}}^2
\over m_{\pi}^2} g^\pi_0 g^\pi_1
\approx 36.83~{\rm GeV}^{-2} \;, \\
{V}^{\prime\rho}_{\rm CSB}(q^2=0) &=&
{g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
f_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
\over m_{\rho}^2} f^\rho_0
\approx 1.07~{\rm GeV}^{-2} \;,
\end{eqnarray}
\end{mathletters}
The one-pion exchange contribution dominates that of the rho;
this is driven by the large $\rho$-$\pi$ mass difference --- recall
$m_{\rho}^{2}/m_{\pi}^{2}\approx 30$. Note that the inclusion
of the rho meson leads to a reduction of the tensor and an
enhancement of the spin-spin components of the pion-exchange
potential. Unlike the vector meson case, we cannot readily
compute the $\pi$-$\eta$ mixing amplitude at $q^2=0$ in our model.
That is, in the pion case, there is no conserved current, so that
the $q^2=0$ mixing can be nonzero~\cite{piekar93,maltman93,chan95}.
Nevertheless, we can compare the results of Eq.~(\ref{zeroq2pieta})
with the ``usual'' $\pi$-$\eta$ mixing potential:
\begin{equation}
{V}^{\pi\eta}_{\rm CSB}(q^2=0) = -
{g_{\lower 2pt \hbox{$\scriptstyle NN\pi$}}
g_{\lower 2pt \hbox{$\scriptstyle NN\eta$}}
\over m_{\pi}^2 m_{\eta}^2}
\langle \pi | H | \eta \rangle \Big|_{q^2=m_\eta^2}
\approx 52.01~{\rm GeV}^{-2} \;,
\label{pietamix}
\end{equation}
where we have input the $\pi$-$\eta$ mixing matrix element evaluated
at its on-shell point, $\langle\pi|H|\eta\rangle|_{q^2=m_\eta^2}=
-4200$~MeV$^{2}$~\cite{coon82}. The contribution from one-pion exchange
is comparable to that from $\pi$-$\eta$ mixing. The $\pi$-$\eta$ mixing
potential may seem slightly larger, but the $NN\eta$ coupling is
ill-determined from two-nucleon data. Indeed, it is believed that the
Bonn potential overestimates it --- a current analysis based on
$\eta$-photoproduction data suggests couplings as low as
$g^2_{NN\eta}/4\pi \alt 0.5$~\cite{tiator94} (see also
Ref~\cite{piekar93}).
The CSB potentials from one-pion exchange have been computed previously
in a nucleon model~\cite{cheung80}. Here the neutron-proton mass
difference, $\Delta M$, generates the breaking. In the specific case
of the class III contribution coming from neutral pion exchange, the
scale of the breaking is set by $\Delta M/2M$. Thus, the isospin breaking
in the quark model is substantially larger than in the nucleon model,
i.e.,
\begin{equation}
\left({3 \over 10}{\Delta m \over m}\right) \bigg/
\left({1 \over 2} {\Delta M \over M}\right) \approx 6 \;,
\end{equation}
so that any CSB observable receiving an important contribution
from $\pi$-$\eta$ mixing will also be affected by the exchange
of neutral pions. The breaking we calculate in the $NN\pi^0$ vertex
is identical to the result of Mitra and Ross~\cite{miller90,mitra67}.
Note that the exchange of charged pions --- and rhos ---
generates a class IV potential which is important in the analysis
of $\Delta A$~\cite{willia87,miller90}. In the charged meson
case, however, the relation between the isospin-violating couplings
in the two models is not simple: it depends on the quark momentum
distribution. Yet, under reasonable assumptions, both sorts of models
seem to generate class IV potentials of comparable
strength~\cite{miller90}.
\section{Results}
\label{secresults}
In this section we compute the CSB potentials for a range of
spacelike momenta. We shall concentrate on class IV contributions
exclusively as we are interested in computing the impact of
the new isospin-violating sources on $\Delta A$. The knowledge of
the $q^2=0$ couplings now no longer suffices. One is forced to model
the momentum dependence of the coupling constants --- including that
of the isospin-violating components. Here we consider two different
estimates for the $q^2$ dependence of the CSB potentials. First, we
simply adopt the momentum dependence which emerges from fits to the
isospin-conserving two-nucleon data. Thus, the ratio of the
isospin-violating to the isospin-conserving coupling, e.g.,
$f^{\omega}_{1}/g^{\omega}_{0}$, remains unchanged. Note that
in the Bonn model $f^{\rho}_{1}/g^{\rho}_{1}$ is also a constant.
We implement this choice by modifying the meson-nucleon ``point''
couplings indicated in Eq.~(\ref{vert}) as per the Bonn B potential
parameters, see Table~\ref{tablethree}. That is,
\begin{mathletters}
\label{bonnff}
\begin{eqnarray}
g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}} \rightarrow
g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}({\bf q}^2) &=&
g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}
(1+{\bf q}^2/\Lambda^2_{\omega})^{-2} \;, \\
g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}} \rightarrow
g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}({\bf q}^2) &=&
g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}
(1+{\bf q}^2/\Lambda^2_{\rho})^{-2} \;.
\end{eqnarray}
\end{mathletters}
This is an additional model assumption.
Here we use {\bf q} to denote the three-momentum transfer; we consider
the form factors in the Breit frame, where $q_0=0$ and $q^2=-{\bf q}^2$.
Second, we compute the
${\cal O}({\bf q}^2)$ isospin-breaking in the couplings in the
nonrelativistic quark model, in order to gauge the uncertainty
in the momentum dependence of the CSB potentials. Let us examine
the isospin-breaking in the Sachs-Walecka form factors~\cite{sachswal},
separated into contributions from the isoscalar or isovector quark
charges. These quantities are related to the $\omega$ and $\rho$
couplings by virtue of our vector dominance assumption. As previously,
we will discuss merely the isospin breaking in the $NN\omega$ vertex in
detail. Now
\begin{mathletters}
\begin{eqnarray}
G_{E,p}^{\omega} &=&
2 g^{+}_u \langle u \rangle_p + g^{+}_d \langle d \rangle_p \;, \\
G_{E,n}^{\omega} &=&
2 g^{+}_d \langle d \rangle_n + g^{+}_u \langle u \rangle_n \;,
\end{eqnarray}
\label{gepm}
\end{mathletters}
and
\begin{mathletters}
\begin{eqnarray}
{G_{M,p}^{\omega} \over 2 M_p} &=&
{1 \over 18}\left( {4\over m_u} \langle u \rangle_p
- {1\over m_d} \langle d \rangle_p \right)
\;, \\
{G_{M,n}^{\omega} \over 2 M_n} &=&
{1 \over 18}\left( {4\over m_d} \langle d \rangle_n
- {1\over m_u} \langle u \rangle_n \right)
\; .
\end{eqnarray}
\label{gmpm}
\end{mathletters}
These expressions are generalizations of Eqs.~(\ref{model1}) and
(\ref{model2}). We have used the notation of Eq.~(\ref{themodel})
in denoting the isoscalar and isovector quark charges and have
introduced $\langle u \rangle_p$, for example, to represent the
Fourier transform of the proton wave function with respect to the
up quark coordinate. We compute the latter in the harmonic
oscillator quark model for simplicity. In the harmonic oscillator
quark model~\cite{dmitra95,bhaduri} the nucleon possesses a mass
$M_N = 2m_1 + m_2$, so that for the proton $m_1=m_u$ and $m_2=m_d$.
For convenience one defines $R^{-2}_{\rho}=\sqrt{3k m_{\rho}}$ and
$R^{-2}_{\lambda}=\sqrt{3k m_{\lambda}}$, where
$m_{\lambda} = 3m_1m_2 / (2m_1 + m_2)$, $m_{\rho} = m_1$, and
$k$ is the spring constant. One finds that~\cite{dmitra95}
\begin{mathletters}
\begin{eqnarray}
\langle u \rangle_p &\equiv& \langle
\exp({\it i} {\bf q}\cdot{\bf r}_u) \rangle_p = 1 - {{\bf q}^2 \over 8}
\left( R_{\rho p}^2 + 3\left(m_d \over M_p\right)^2 R_{\lambda p}^2 \right)
+ {\cal O}({\bf q}^4) \;, \\
\langle d \rangle_p &\equiv& \langle
\exp({\it i} {\bf q}\cdot{\bf r}_d) \rangle_p = 1 - {3{\bf q}^2 \over 2}
\left(m_u \over M_p\right)^2 R_{\lambda p}^2 + {\cal O}({\bf q}^4) \;, \\
\langle u \rangle_n &\equiv& \langle
\exp({\it i} {\bf q}\cdot{\bf r}_u) \rangle_n = 1 - {3{\bf q}^2 \over 2}
\left(m_d \over M_n\right)^2 R_{\lambda n}^2 + {\cal O}({\bf q}^4) \;, \\
\langle d \rangle_n &\equiv& \langle
\exp({\it i} {\bf q}\cdot{\bf r}_d) \rangle_n = 1 - {{\bf q}^2 \over 8}
\left( R_{\rho n}^2 + 3\left(m_u \over M_n\right)^2 R_{\lambda n}^2 \right)
+ {\cal O}({\bf q}^4) \;.
\end{eqnarray}
\label{qformf}
\end{mathletters}
We write the Fourier transforms in Eq.~(\ref{qformf}) through
${\cal O}({\bf q}^2)$ only. This suffices to make contact with
the hadronic form factors. Moreover, one cannot expect the
nonrelativistic quark model to be reliable at still larger momentum
transfers. We must now relate the above electric and magnetic form
factors to the $f$'s and $g$'s present in the definition of the vertex,
Eq.~(\ref{vert}). Following the usual relation between the electromagnetic
form factors $G_E$, $G_M$ and $F_1$, $F_2$, vector dominance dictates that
\begin{mathletters}
\label{vectdom}
\begin{eqnarray}
G_{E,N}^{\omega}(q^2) &=& g_N^{\omega}(q^2)
+ {q^2 \over 4 M_N^2} f_N^{\omega}(q^2) \;,
\\
G_{M,N}^{\omega}(q^2) &=& g_N^{\omega}(q^2) + f_N^{\omega}(q^2)
\;.
\end{eqnarray}
\end{mathletters}
Fits to the electromagnetic form factor data indicate that
$F_1(q^2)$ and $F_2(q^2)$ fall with different rates in $q^2$;
vector dominance implies that this should be true of
$g_N^{\omega,\rho}$ and $f_N^{\omega,\rho}$ as well. Note, this
is at odds with the Bonn model, as it assumes that the ratio
$f_N^{\omega,\rho}/g_N^{\omega,\rho}$ is constant.
We proceed as follows. We compute $f_N^{\omega}$ and $g_N^{\omega}$
to ${\cal O}({\bf q}^4,{\Delta m}^2)$, using
Eqs.~(\ref{gepm}-\ref{vectdom}). Then we estimate the ``effective''
$\Lambda_{\omega}$, as defined in Eq.~(\ref{bonnff}), required to
reproduce the isospin breaking computed to
${\cal O}({\bf q}^4,{\Delta m}^2)$ and use that
$\Lambda_{\omega}$ in our subsequent computation of
the spin-singlet--triplet mixing angles. Thus,
\begin{mathletters}
\begin{eqnarray}
g_{0}^{\omega} + g_{1}^{\omega}\tau_z &=& \left[
1 - {{\bf q}^2 R^2 \over 6} \right] +
{\Delta m \over m} \left[ {5{\bf q}^2 \over 24 M^2} -
{{\bf q}^2 R^2 \over 72} \right] \tau_z +
{\cal O}({\bf q}^4,{\Delta m}^2) \;, \\
f_{0}^{\omega} + f_{1}^{\omega}\tau_z &=&
{5\over 6}{\Delta m \over m}
\left[ 1 - {{\bf q}^2 R^2 \over 3} -
{{\bf q}^2 \over 4 M^2} \right] \tau_z +
{\cal O}({\bf q}^4,{\Delta m}^2) \;.
\label{breako}
\end{eqnarray}
\end{mathletters}
Several remarks are in order. First, note that we have defined
$R^{-2}=\sqrt{3k m}$, where $m$ is the average mass of the up
and down quarks. From Eq.~(\ref{breako}) we observe
$f_0^{\omega}=0$ to ${\cal O}({\bf q}^4,\Delta m^2)$; this is
consistent with the Bonn model, which assumes $f_0^{\omega}=0$
for all $q^2$. We have performed the same calculations for
the $NN\rho$ vertex as well. In this case, one finds results at
odds with the Bonn model, as $f_1^{\rho}/g_1^{\rho}$ is not constant
to ${\cal O}({\bf q}^4,\Delta m^2)$. Note that at nonzero ${\bf q}^2$
CSB potentials beyond those enumerated in
Eqs.~(\ref{vomegab})-(\ref{vpionb}) may exist. For example, at
${\cal O}({\bf q}^2,\Delta m)$ a new CSB contribution arises
from the combination $f_1^{\rho}g_0^{\rho}$. Yet, like the rho
contribution to the CSB potential given in Eq.~(\ref{vrhob}), it
is not numerically important, due to the small value of
$g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}$ in the Bonn
model --- recall that $g_{\lower 2pt \hbox{$\scriptstyle NN\omega$}}^2 /
g_{\lower 2pt \hbox{$\scriptstyle NN\rho$}}^2 \approx 27$~\cite{machl89}.
Let us proceed to examine the impact of Eq.~(\ref{breako}) on the omega
contribution to the class IV CSB potential. We fix the scale $R$ by
requiring that the isospin conserving vertex, $g_0^{\omega}$, fall in
${\bf q}^2$ at the rate given by the Bonn model, so that
$R=\sqrt{12}/\Lambda_{\omega} \approx .37$~fm. We choose $R$ in this
manner as our primary interest is in determining the fall-off of the
isospin-breaking potential {\it relative} to the isospin-conserving one.
Noting Eq.~(\ref{vomegab}), we consider
\begin{mathletters}
\begin{eqnarray}
f_{1}^{\omega} g_{0}^{\omega} &=&
{5 \over 6} {\Delta m \over m}
\left( 1 - 6{{\bf q}^2 \over \Lambda_{\omega}^2} -
{{\bf q}^2 \over 4 M^2} \right) +
{\cal O}({\bf q}^4,{\Delta m}^2) \\
\quad &\equiv& {5 \over 6} {\Delta m \over m}
\left( 1 - 4{{\bf q}^2 \over {\widetilde\Lambda}_{\omega}^2} +
{\cal O}({\bf q}^4) \right) \;.
\end{eqnarray}
\end{mathletters}
By replacing the $\Lambda_{\omega}$ of Eq.~(\ref{bonnff}) with
the ${\widetilde\Lambda}_{\omega}$ given above, such that
\begin{equation}
{\widetilde\Lambda}_{\omega}^2 = \Lambda_{\omega}^2
\left( { 4 \over 6 + \Lambda_{\omega}^2 /4 M^2 } \right) \;,
\label{tildel}
\end{equation}
we obtain an expression for the CSB potential, Eq.~(\ref{vomegab}),
which is of the form given by our original prescription, Eq.~(\ref{bonnff}),
yet is equivalent to the isospin-breaking calculated in the harmonic
oscillator quark model at ${\cal O}({\bf q}^2,\Delta m)$. Numerically,
the Bonn model $\Lambda_{\omega}=1850$ MeV is changed to
$\widetilde{\Lambda}_{\omega}=1401$ MeV. At this order
the coefficient of $g_1^{\rho}$ is not negative, so that we cannot carry
out the above exercise for the rho as well. The rho's numerical impact
on the CSB potential is small, so that this gap does not impact our
uncertainty estimate in any significant way. We will proceed to compute
the spin-singlet--triplet mixing angles for the potential given by
Eqs.~(\ref{vomegab}) and (\ref{bonnff}) for both the $\Lambda_{\omega}$
of the Bonn potential and the $\widetilde{\Lambda}_{\omega}$ of
Eq.~(\ref{tildel}).
In Fig.~\ref{figone} we present estimates of the CSB potentials given
in Eqs.~(\ref{vomegab}), (\ref{vrhob}), and (\ref{vrhoomegab}) using
Eq.~(\ref{bonnff}) with the $\Lambda_{\omega,\rho}$ of the Bonn model.
The qualitative conclusions we draw here are not sensitive to the choice
of $\Lambda_{\omega,\rho}$, so that we simply present the potentials
computed in the Bonn model. The solid line is the CSB potential which
results from $\rho$-$\omega$ mixing, Eq.~(\ref{vrhoomegab}), if
the on-shell value of the $\rho$-$\omega$ mixing amplitude is employed
for the entire range of momenta. This is the potential traditionally
used in studies of CSB observables. A potential of this strength is
required to describe the analyzing power difference $\Delta A$
measured in elastic $\vec{n}-\vec{p}$ scattering~\cite{holz87}. In
contrast, the dashed line results if the momentum-dependent
$\rho$-$\omega$ mixing amplitude of Ref.~\cite{piewil93} is employed
in Eq.~(\ref{vrhoomegab}) --- this is too small to
fit the data~\cite{iqnisk94}, yet
a model in which the vector mesons couple to conserved currents
must yield a vanishing mixing amplitude at $q^2=0$~\cite{oconn94}.
We have not extended our model to describe $\rho$-$\omega$ mixing;
the vector dominance assumption we use implies, however, that the
$q^2=0$ mixing must be zero in this framework. We take the
momentum-dependence of the mixing amplitude computed by
Piekarewicz and Williams~\cite{piewil93} as archetypal. This latter
CSB potential in itself would upset the previous agreement with
experiment. However, the new sources of isospin violation computed
here are sufficient to restore the agreement. In particular, the
contribution from omega-meson exchange, given by the dash-dotted line,
is large and comparable in magnitude to the one arising from on-shell
$\rho$-$\omega$ mixing. We have also computed the contribution from
the rho-meson, given by the dotted line, though it is negligible due
to the small $NN\rho$ vector coupling.
In Fig.~\ref{figtwo} we display the above CSB potentials in
configuration space. The potentials have been normalized so
that the areas under the curves equal $V(q^2=0)$. Qualitatively,
the trends observed in Fig.~\ref{figone} remain: we obtain large
contributions from on-shell $\rho$-$\omega$ mixing and omega-meson
exchange and small corrections to the latter from off-shell
$\rho$-$\omega$ mixing and rho-meson exchange. These results are
suggestive, yet we can obtain a precise estimate of the impact of
the enumerated isospin-violating sources on $\Delta A$ by calculating
the spin-singlet--triplet mixing angles,
$\bar{\gamma}_{\lower 2pt \hbox{$\scriptstyle J$}}$.
These are the dynamical quantities driving
$\Delta A$~\cite{willia87,miller90}.
Recall that the elastic scattering amplitude of two spin-$1/2$
particles is specified by six invariant amplitudes
$a, b, c, d, e,$ and $f$~\cite{miller90}, so that
\begin{eqnarray}
\widehat{M}
&=&{1\over 2} \Big[ (a + b) +
(a-b) ({\bf \sigma}_1\cdot \hat{\bf n}) ({\bf \sigma}_2\cdot \hat{\bf n})
\nonumber \\
&+& (c+d) ({\bf \sigma}_1\cdot \hat{\bf m}) ({\bf \sigma}_2\cdot \hat{\bf m})
+ (c-d) ({\bf \sigma}_1\cdot \hat{\bf l}) ({\bf \sigma}_2\cdot \hat{\bf l}) \\
&+& e ({\bf \sigma}_1 + {\bf \sigma}_2)\cdot \hat{\bf n}
+ f ({\bf \sigma}_1 - {\bf \sigma}_2)\cdot \hat{\bf n}
\Big]
\nonumber \;,
\end{eqnarray}
where
\begin{equation}
\hat{\bf l}\equiv { {\bf k}_f + {\bf k}_i \over
{| {\bf k}_f + {\bf k}_i |}} \;, \quad
\hat{\bf m}\equiv { {\bf k}_f - {\bf k}_i \over
{| {\bf k}_f - {\bf k}_i |}} \;, \quad
\hat{\bf n}\equiv { {\bf k}_i \times {\bf k}_f \over
{| {\bf k}_i \times {\bf k}_f |}} \;,
\end{equation}
and ${\bf k}_i$ and ${\bf k}_f$ are the initial and final
c.m. momenta of particle 1. The $\vec{n}-\vec{p}$ analyzing power difference
is nonzero only if accompanied by spin-singlet--triplet mixing,
specifically
\begin{equation}
\Delta A(\theta) \equiv A_n(\theta) - A_p(\theta) =
2 {\rm Re } (b^* f)/\sigma_0 \;,
\end{equation}
where $\sigma_0$ is the unpolarized differential cross section.
The spin-singlet--triplet mixing is controlled by $f$. Neglecting
electromagnetic effects, $f$ is connected to the mixing angles
$\bar{\gamma}_{\lower 2pt \hbox{$\scriptstyle J$}}$ via~\cite{gersten81}
\begin{equation}
f(k,\theta) ={ i \over 2k} \sum_{J=1}^{\infty}
(2J +1) \sin (2 \bar{\gamma}_{\lower 2pt \hbox{$\scriptstyle J$}})
\exp ( i \bar{\delta}_{\lower 2pt \hbox{$\scriptstyle J$}} +
i \bar{\delta}_{\lower 2pt \hbox{$\scriptstyle JJ$}} )
d_{10}^J (\theta) \;,
\end{equation}
where the $d_{10}^J(\theta)$ are Wigner functions and the
$ \bar{\delta}_{\lower 2pt \hbox{$\scriptstyle J$}}$ and
$ \bar{\delta}_{\lower 2pt \hbox{$\scriptstyle JJ$}}$
are the singlet
and uncoupled triplet bar phase shifts, respectively.
In a distorted-wave Born approximation the mixing angles
themselves are given by~\cite{willia87}
\begin{equation}
\bar{\gamma}_{\lower 2pt \hbox{$\scriptstyle J$}}
= -4Mk \sqrt{J(J+1)}
\int_0^\infty dr r^{2} R_{J}(r) V_{IV}(r) R_{JJ}(r)
\equiv \int_0^\infty dr
I_{{\lower 2pt \hbox{$\scriptstyle J$}}}(r)
\label{mixangle} \;,
\end{equation}
where we have introduced the class IV CSB potential
\begin{equation}
V_{IV}(r) \equiv {1 \over 2M^{2}r}
{dV_{\rm CSB}(r) \over dr} \;.
\end{equation}
Note that
$R_{J}(r)$ and $R_{JJ}(r)$ are the spin-singlet and triplet
radial wave functions, respectively, for $NN$ scattering
in the $L=J$ channel. Distortion effects are incorporated
through these radial wave functions; we assume them adequately
described by solutions to the Reid soft-core
potential~\cite{reid68}. In Table~\ref{tablefour} the first
four nonvanishing mixing angles, $J=1 - 4$, are presented at a
laboratory energy of 183 MeV. In addition, the integrand from
which $\bar{\gamma}_{1}$ is obtained, that is, $I_{1}(r) $
in Eq.~(\ref{mixangle}), is plotted in Fig.~\ref{figthree}.
This represents the class IV potential suitably weighted by
realistic $NN$ wave functions. Three calculations are presented
for comparison. The solid line is obtained using Eq.~(\ref{vrhoomegab})
and the on-shell value of the $\rho$-$\omega$ mixing amplitude; the
area under this curve is the mixing angle required to reproduce
the $\Delta A$ data. In the dashed line we have combined the
off-shell $\rho$-$\omega$ mixing contribution described above with
the isospin-violating vertex contributions arising from omega- and
rho-meson exchange. Albeit form factor uncertainties in the
isospin-violating vertices, this is our best estimate of the
mixing angle contribution. The vertices in this figure were
evaluated using Eq.~(\ref{bonnff}) and the Bonn cutoff parameters
$\Lambda_{\omega,\rho}$ tabulated in Table~\ref{tablethree}.
We have also combined the {\it on-shell} $\rho$-$\omega$ mixing
contribution with the above vertex contributions, even if our model
is not consistent with a nonzero mixing amplitude at $q^2=0$ --- this
is shown by the dash-dotted line. The integrand in this case is
considerably larger than the other two estimates. The $J=1$ mixing
angles for these integrands are displayed in parentheses next to the
curve labels. The agreement between the first two calculations
is very good. Indeed, the contribution from omega-meson exchange,
together with small corrections from off-shell $\rho$-$\omega$ mixing
and rho-meson exchange, results in a 3\% reduction in the value of
$\bar{\gamma}_{1}$, relative to the on-shell value. This kind of
agreement --- at the few percent level --- is maintained throughout
all the examined partial waves, note Table~\ref{tablefour}. These
computations have also been performed with the form factor
$\widetilde\Lambda_{\omega}$, Eq.~(\ref{tildel}), estimated in the
harmonic oscillator quark model. The mixing angles obtained in this
fashion vary by about 10\% in the important partial waves from those
computed with the Bonn form factors; note that
$\bar{\gamma}_{1}=.036^{\circ}$, rather than $.033^{\circ}$. For
a detailed comparison, see Table~\ref{tablefour} --- the mixing angles
which use the harmonic oscillator quark model results to estimate the
${\bf q}^2$ dependence of the CSB potentials are shown in parentheses.
The mixing angles computed with isospin-breaking meson-nucleon vertices
and off-shell $\rho$-$\omega$ mixing in the two approaches bracket the
old on-shell $\rho$-$\omega$ mixing results for $J=1-3$, so that these
new estimates are also quite close to the ``old'' on-shell results.
$\widetilde \Lambda_{\omega}$ is some 3/4 of $\Lambda_{\omega}$, yet
the above calculations show that the $\Delta A$ at 183 MeV is essentially
dominated by the $q^2=0$ physics. Note that if one were to assume a
momentum-independent $\rho$-$\omega$ mixing amplitude~\cite{miller94} and
to include the contributions from omega and rho exchange an increase of
almost a factor of two relative to the above mixing angle estimates would
result.
\section{Conclusions}
\label{secconcl}
We have studied the charge-symmetry breaking in the $NN$
potential arising from isospin-violating meson-nucleon coupling
constants. The isospin-violating couplings are obtained by
computing matrix elements of quark currents of the appropriate
Lorentz and flavor structure between nucleon states. We have used
a nonrelativistic quark model to evaluate these matrix elements,
yet our estimates at $q^2=0$ depend merely on the spin and flavor
structure of the nucleon wave function, rather than on the details
of the quark momentum distribution. Thus, in the vector meson sector,
for example, our model estimates at $q^2=0$ depend on our vector
dominance assumption, but little else. We have also studied isospin
breaking in the $NN\pi$ and $NN\sigma$ vertices. No isospin-violations
exist in the $\sigma$ vertex at $q^2=0$. We have found that the breaking
in the $NN\pi$ vertex depends on whether the $\pi$N coupling is of
pseudoscalar or pseudovector character --- no isospin breaking results
if pseudovector coupling is assumed. However, a pseudoscalar $\pi$N
coupling is commonly used in studies of CSB, and the breaking we find
in the vertex is substantially larger than the breaking computed in
hadronic models of neutral pion exchange. Thus, any CSB observable
receiving an important contribution from $\pi$-$\eta$ mixing will also
be affected by the exchange of neutral pions.
We have found that omega-meson exchange is an important component of the
class IV charge-symmetry-breaking $NN$ potential needed to describe the
analyzing power difference measured in elastic $\vec{n}-\vec{p}$ scattering
at low energies. The potential which emerges from the isospin-violating
$NN\omega$ vertex is identical in structure to that from $\rho$-$\omega$
mixing~\cite{ghp95}. Moreover, our estimates indicate that these two
contributions --- with the mixing amplitude fixed at its on-shell value
--- are comparable in magnitude and identical in sign at $q^2=0$. Models
in which the vector mesons couple to conserved currents, of which ours
is an example, have no $\rho$-$\omega$ mixing at $q^2=0$~\cite{oconn94}.
We have found that isospin-violation in the $NN\omega$ vertex can generate
a CSB potential of sufficient magnitude to fill the phenomenological role
required by the IUCF measurement of $\Delta A$ at 183 MeV.
The isospin-violating couplings we have computed at $q^2=0$ do not suffice
to make a quantitative prediction of the CSB potential needed for the IUCF
experiment. One must compute the $q^2$ dependence of the $NN$~meson vertex
as well --- including that of the isospin-violating pieces. We have
considered two simple estimates. The first is simply a prescription:
we modify the ``point'' couplings by introducing hadronic form
factors according to the Bonn B potential. This assumes that the
relative strength of the isospin-breaking potential found at $q^2=0$
persists at finite $q^2$ as well. In the second we compute the isospin
breaking to ${\cal O}({\bf q}^4,\Delta m^2)$ using the spatial wave
functions of the harmonic oscillator quark model and find the hadronic
form factor for omega exchange needed to reproduce the isospin breaking
computed to the above order. The use of the spatial component of the
nucleon wave function is required here, so that this estimate is rather
more model dependent than our $q^2=0$ results. We find that the use of
the latter estimate yields slightly larger CSB potentials.
Armed with estimates of the momentum dependence of the $NN$~meson
vertex, we have computed the spin-singlet--triplet mixing angles
$\bar{\gamma}_{\lower 2pt \hbox{$\scriptstyle J$}}$: these are the
fundamental dynamical quantities driving $\Delta A$.
Our $\bar{\gamma}_{\lower 2pt \hbox{$\scriptstyle J$}}$ computation
is realistic as we have used the Reid soft-core potential to generate
the distortions in the $NN$ wave function. We have computed the
spin-singlet--triplet angles using three different sources of isospin
violation: (1) $\rho$-$\omega$ mixing with the amplitude fixed at its
on-shell value, (2) off-shell $\rho$-$\omega$ mixing plus omega- and
rho-meson exchange, and (3) on-shell $\rho$-$\omega$ mixing plus omega-
and rho-meson exchange. The first case, used in the original estimates
of $\Delta A$, represents a ``baseline'' value, as it fits the data.
A CSB potential of this magnitude accounts for a substantial fraction
of the measured value of $\Delta A$ at 183 MeV. The second case, which
should be regarded as our best estimate, yields values for
$\bar{\gamma}_{\lower 2pt \hbox{$\scriptstyle J$}}$ that are
within 10\% of those obtained with on-shell $\rho$-$\omega$ mixing,
for the important partial waves. In contrast, case (3) results in a
factor-of-two enhancement relative to the original calculation using
on-shell $\rho$-$\omega$ mixing. Two important results thus emerge from
the present work. First, we have found a new source of isospin violation,
namely in the $NN\omega$ vertex, which can fill the role demanded by the
data. Second, we have shown that insisting upon a $\rho$-$\omega$ mixing
amplitude held constant at its on-shell value, after including the
contribution from omega-meson exchange, results in a class IV potential
too large to be consistent with the IUCF $\Delta A$ measurement.
\acknowledgments
We thank V. Dmitra\v{s}inovi\'c and S.J. Pollock for fruitful
discussions, A. Thomas for a helpful
suggestion, and S. Capstick for useful
conversations.
This work was supported by the DOE under
Contracts Nos. DE-FG02-87ER40365 (S.G. and C.J.H.),
DE-FC05-85ER250000 (J.P.), and DE-FG05-92ER40750 (J.P.).
|
\section{Introduction}
Anisotropies in the Cosmic Microwave Background Radiation (CMBR)
contain a wealth of information about the conditions and processes
that led to the formation of large scale structures in the universe.
Whether the correct model of the evolution of the universe involves
baryons, exotic cold dark matter,
massive neutrinos, topological defects,
a standard ionization history or reionization at a modest
redshift, the content and history
of the universe affect the observed anisotropies of the
CMBR (see \cite{White} for a review).
Constrained by the measured level of anisotropy on large angular
scales~(\cite{DMR2,FIRS}), different models predict varying levels of
anisotropy on smaller scales.
Several observing teams~(\cite{MSAMI,ARGO,MAX,UCSB,SK}) including our own
have reported detections of anisotropy on angular scales near $1^\circ$.
Improvements in these measurements should
help discriminate between the current models of
large scale structure formation.
The Python instrument is designed to search for anisotropies
at an angular scale near $1^\circ$. The first observations (hereafter PyI)
with Python
were made between 1~and~15 January 1993 at the Amundsen-Scott South Pole
station~(\cite{PyI}, hereafter Paper~I).
Statistically significant signals were detected in observations
at high Galactic latitude
($|b| > 49^\circ$).
We report here the results of a second set of
observations (hereafter PyII),
made between 12 and 23 December 1993 from the same site.
The bulk of the observing time in the second season was spent
on the PyI field (``Field~A"), in an effort to repeat that measurement.
The remainder was spent on a set of interleaving spots (``Field~B").
The observation of the Field~A spots serves as a further
check for the effects of interfering signals that would differ
from one year to the next,
including effects from atmospheric emission, cosmic ray hits in
the bolometers, and radiofrequency interference.
Additionally, the Sun was approximately $20^\circ$ further away from the
observed fields during the PyII season than it was for PyI.
\section{Instrument}
The Python instrument~(Paper~I, \cite{Ruhlthesis}) consists of a 0.75~m
diameter off-axis parabolic
telescope that couples radiation into
a $2 \times 2$ array of bolometric detectors.
The beam response is well approximated by a Gaussian with
a full width at half maximum of $(0.75 \pm 0.05)^\circ$.
Radiation from the sky is first reflected off a vertical flat before
being focused into the cryostat by the primary.
Rotation of the flat about a vertical axis moves the detector response
horizontally across the sky.
The telescope is mounted on an
azimuth-elevation mount and is
surrounded by a large shield that protects the
instrument from being illuminated by the Sun or Earth.
Inside the cryostat there are four corrugated feed horns
at the focal plane
of the primary. Radiation that enters a feed horn passes
through a set of single-mode waveguide filters
that define the passband at $\lambda = 3.3$~mm
before reaching the bolometric detectors.
The detectors use a layer of bismuth evaporated on a diamond wafer
as the absorbing element, and a chip of neutron transmutation doped
germanium as the thermistor element. They are operated at
50~mK, cooled by a $^3$He guarded
adiabatic demagnetization refrigerator~(\cite{R_and_D}).
An additional bolometer (the ``dark channel"),
mounted on the cold stage but kept in a
sealed cavity, acts as a monitor for extraneous pickup.
New bolometers, constructed using the same methods as the
originals,
were installed for the PyII observing season.
However, two of the four new optical channels did not work well.
One channel was a factor of 3 less sensitive than typical
(6~mK$\cdot \sqrt{\mbox{s}}$ instead of 2~mK$\cdot \sqrt{\mbox{s}}$);
the other was unusable due to an electrical problem in the dewar.
Fortunately, one good channel was on the upper row, and the other was on
the lower, making possible the confirmation of
the bulk of the PyI data.
The two most significant changes to the telescope between PyI and PyII
were the improved balancing of the chopping mechanism for the
external switching flat, and the installation of a microwave-absorbing
guard ring around the primary.
\section{Observations}
The Sun, Moon, and two sources in the Carinae nebula were used
as absolute pointing references each
season; additionally, Venus was used during the PyI season.
We estimate our absolute pointing accuracy to be $\pm 0.1^\circ$
each year, or $\pm 0.15^\circ$ for the relative pointing
accuracy of the PyI and PyII datasets.
The same observing strategy was used for PyI and PyII, and is
described in Paper~I. The combination of a fast 3-beam chop
(2.5~Hz full cycle) and slower beamswitching (10 to 30 seconds per position)
yields a 4-beam response on the sky,
\begin{equation}
\Delta T_{j}^k = -\frac{1}{4} T_j + \frac{3}{4} T_{j+1}
-\frac{3}{4} T_{j+2} + \frac{1}{4} T_{j+3},
\end{equation}
where the $T_j$'s are the antenna temperatures of patches on
the sky separated by $2.75^\circ$ along a horizontal line.
A scan consists of measuring $\Delta T_j^k$ three times
successively~($k=1,2,3$) at each of 7 positions $j$
on the sky. The left and right hand channels in a given row of the array
measure many of the same 4-beam patterns; the spots measured by
the left hand side are given by $j=1,...,7$, while those made by the
right correspond to $j=2,...,8$.
The time-ordering of the right and left-hand
3-beam measurements (``stares'') are
reversed from one value of $j$ to the next. For the first 4-beam pattern
($j=1$ for the left channels, $j=2$ for the right),
the left stare is measured first. For the second
4-beam pattern the right stare is measured first,
and so on.
This causes a drifting 3-beam offset to appear as an
oscillating 4-beam signal on the sky.
However, the measured linear drifts in the 3-beam offset
within each scan are both small and symmetric about zero, making
this effect unimportant.
The average 3-beam offsets are less than 1~mK for
all the PyI and PyII channels, and the average slopes are
within $2\sigma$ of zero, with an error of $\sigma \sim 1 \mu$K/stare.
\section{Calibration}
The PyI observations were calibrated by a combination
of elevation scans and the placement of known-temperature
blackbody and low emissivity foam loads
in the optical path, as described in Paper~I and \cite{R_capri}.
The PyII observations were calibrated using the foam loads;
the loads were in turn calibrated by placing a
large liquid-nitrogen-cooled load beneath the
dewar and alternately switching the foam and two known-temperature loads into
the optical path. This calibration
of the foam loads was done {\it in situ} on the telescope during
the PyII season,
rather than in a 300~K room, reducing possible systematic effects
in the measurement. The DC gain measured
using the foam loads was converted to an AC gain
by comparing the signals seen while switching on and off a foam
load slowly (0.1~Hz) with those seen while switching at the frequency used
for the observations (5~Hz).
The relative calibrations of the various bolometer
channels (four channels for PyI, three for PyII)
were checked using two sources in
the Carinae nebula; the results from all seven channels
lie within 15\% of the average.
The two sources in the Carinae nebula are at
($\alpha$,$\delta$) = $(10^{\mbox{\scriptsize h}}44^{\mbox{\scriptsize m}},
-59.64^\circ)$ and
$(10^{\mbox{\scriptsize h}}33^{\mbox{\scriptsize m}},-57.95^\circ)$
in J1994 coordinates.
The mean signals from them are
($9.0 \pm 0.3$) and ($4.2 \pm 0.1$)~mK in our beam, respectively,
with the errors representing only the statistical uncertainty.
The first source was also used as a pointing reference.
The statistical error on the determination of the average gain
is 5\%. Our estimates of possible systematic
effects are at the level of $<20$\%. We therefore adopt a
gain uncertainty of 20\%.
\section{Analysis}
The data analysis presented here differs slightly from
that used for the previously
reported PyI results; using the new method does not significantly change
those results. We describe here the new method, which is used to arrive at
all the results in this paper.
The output of each detector is sampled at 100~Hz
and demodulated in software, synchronously with the
motion of the chopping flat.
Two lockin demodulations are used; one (the ``optical phase") is
maximally sensitive to the 3-beam signal from the sky, while the second
(the ``quadrature phase") is shifted by $90^\circ$ from the
optical phase and should have no response to a stationary sky signal.
We confine our discussion to the optical phase
data, giving results from the quadrature phase where relevant.
Within each of the 42 stares in a scan,
the mean and variance
$\sigma_{S_j}^2$ of the lockin values are calculated.
The means are then combined in a pairwise fashion into 4-beam values
$\Delta T_j^k$ as described above, and an average variance
$\sigma_{S}^2$ is calculated from the 42 values of $\sigma_{S_j}^2$.
Thus $\sigma_{S}^2$ is a measure of the noise which contains drifts
only within 10 to 30 second long stares.
As previously described, each channel measures a 4-beam value
$\Delta T_j^k$ three times in succession $(k = 1,2,3)$
for seven sky positions $j$. From these we find
a mean for each channel at each position $j$,
and an average error on those means, $\sigma_m$.
Drifts up to the
time separating observations of successive 4-beam positions
(roughly 3 minutes), are included in $\sigma_m$.
The variance from a celestial signal does not contribute to
$\sigma_{m}$, and $\sigma_{m}$ can therefore be used as an unbiased
statistic for cutting scans.
The noise estimates described above are calculated for each channel
in every complete scan.
We use $\sigma_{S}$ and $\sigma_{m}$, normalized to
a ``stare" duration of 30 seconds,
on a channel by channel basis to remove scans which have
been contaminated by excessive noise.
The first cut removes scans with $\sigma_{S}$
greater than 1.5 times the peak of
its distribution. This procedure removes 20 to 25\% of the
scans for each channel in PyI, and 5 to 10\% of those in PyII.
The second cut removes scans with high values of $\sigma_{m}$;
the value at which scans are cut is placed so as to minimize the
average final errorbar in the binned data.
This step removes 19 to 27\% of the scans for PyI, and
4 to 5\% of those in PyII. The cuts remove more PyI data than PyII
because of the poorer weather during that season.
In all, approximately 50 hours (depending on the channel) of PyI data
passed the cuts. For PyII, roughly 24 hours of data on the overlapping
field (Field~A), and 9 hours on the interleaving field (Field~B)
remained after the cuts.
The average value of $\sigma_{S}$ from all scans that pass the cuts
is a good indicator of chopped detector noise. For PyI, the noise
in the four channels is 2.5, 2.9, 2.1 and 1.9 mK~$\cdot \sqrt{\mbox{s}}$;
For the three operational channels of PyII, it is
2.1, 1.3 and 5.7 mK~$\cdot \sqrt{\mbox{s}}$.
These values are in good agreement with those found
from the quadrature values of both $\sigma_{S}$ and $\sigma_{m}$.
The optical phase value of $\sigma_{m}$ is higher (for the
sensitive channels), due to atmospheric contamination.
For PyI it is 4.1, 4.2, 4.0 and 3.9 mK~$\cdot \sqrt{\mbox{s}}$;
for PyII these values are 2.9, 2.7 and 5.6 mK~$\cdot \sqrt{\mbox{s}}$.
For sky positions observed by more than one channel,
the 4-beam averages from neighboring
channels are combined into a single value within a scan
by forming a weighted average of the
left and right-hand channels. The average value of $1/\sigma_S^2$ over
all uncut scans is used as the weight.
If one of the channels is cut by the previously described
procedures, its neighbor is cut as well.
After this treatment,
a statistical mean and error are calculated from the uncut scans
at each of the sky positions $j$.
The errors are within 10\% of those expected given the distribution
of $\sigma_m$, indicating there is little if any additional
atmospheric contamination.
These means and errors, after multiplying by 1.07 to
correct for the estimated
atmospheric absorbtion in our band, are the final values and errors
for the 4-beam temperature differences on the sky.
\section{Results}
The results of the analysis of the PyI data
and of the overlapping portion (Field~A) of the PyII data set
appear in Figure~\ref{GE_Ts}.
The agreement is good; taking the difference of the two data sets
[ $(\mbox{PyI} - \mbox{PyII})/2$ ]
gives a result that agrees well with zero signal
(reduced $\chi^2 \equiv \chi^2_\nu = 12./15$).
The weighted mean of the two data sets is shown
as the filled circles in Figure~\ref{GEGC_Ts};
here the agreement with zero signal is very poor ($\chi^2_\nu = 191./16$).
The contrast between the $\chi^2$'s for the summed and differenced
data sets indicates that the signal in the two data sets is the same.
A set of patches (Field~B) that interleaves those from Field~A was also
observed during the PyII season. The locations of the
Field~B beams are found by moving one half of a chopper throw
in negative right ascension ($1.38^\circ$ on the sky) from the beams
that make up Field~A.
Analyzing the Field~B data
leads to the values shown as open squares in Figure~\ref{GEGC_Ts}.
Less time was spent observing this field,
and the errors are larger than on Field~A.
The agreement with zero sky signal is
good for Field~B ($\chi^2_\nu = 14./15$).
The results shown in Figure~\ref{GEGC_Ts} were measured in a single
frequency band, so we cannot spectrally discriminate between CMBR
anisotropies and the various possible foregrounds. In Paper~I we
discussed the expected levels of galactic foreground contamination,
which are smaller than the signals seen. Proceeding under the
assumption that these signals represent fluctuations in
the CMBR, we multiply the values plotted in
Figure~\ref{GEGC_Ts} by 1.24 to convert them from
Rayleigh-Jeans temperature units to temperature differences
in a 2.73~K blackbody, and use
the integrated likelihood function to form
confidence interval estimates for the sky signal in two models
with Gaussian CMBR fluctuations. For both analyses we use
the full correlation matrix including off-diagonal elements
describing correlations present because of the theoretical model and
those induced by the observing strategy.
The first model consists of a sky with a Gaussian autocorrelation
function~(GACF) with a coherence angle of $\theta_c = 1^\circ$.
We set limits on $\sqrt{C_0}$, the rms amplitude of the fluctuations in
$\Delta T/T$. The uncertainty in beamwidth leads to less than
a 2\% error in $\sqrt{C_0}$.
The GACF limits can be converted to an estimate
of the flat band power sampled by our window function. Following the recipe
of \cite{Bond1}, we find for our window function,
$\Delta T_{\mbox{\scriptsize \it Band Power}} = 0.73\Delta T_{GACF}$.
The effective center of our window function
for a flat band power spectrum lies at $\ell_{e} = 93$.
We also use a model with an uncorellated sky,
$C(\theta) = \delta(1-\cos\theta)$,
to find the rms of our
data set, from which we derive another band power estimate.
Correlations due to beam overlap are less than 1\%, and we ignore them.
This analysis gives results that are within a few percent of those
found by converting
the GACF limits, indicating that the flat band power is insensitive
to the form of the corellation function used in the likelihood
calculation.
The results for the various combinations of data sets
and analyses are given in Table~\ref{table_res}.
The confidence intervals quoted in the table do not include the
20\% calibration error. The first four
entries in Table~\ref{table_res} show that the
signal seen in the first season appears again in the second.
The last entry in Table~\ref{table_res} gives our final result
using the data from both seasons and both fields.
\section{Conclusions}
After detecting a signal in our first year of observations, we
re-observed the same portion of the sky and detected the same signal.
This provides further evidence that the signal is
celestial rather than systematic in nature.
Additionally, an interleaving set of spots was
observed during the second season.
Likelihood analyses on the entire data set are used to
derive two estimates of the sky signal. First, we find
$79^{+28}_{-19} \mu$K for the total sky rms for a
Gaussian autocorrelation model with a coherence angle of $1^\circ$.
Second, we find a band power of $57^{+20}_{-14} \mu$K,
using an autocorrelation fuction given by
$C(\theta) = \delta(1 - \cos\theta)$.
The stated errors are 1~$\sigma$ limits that include the 20\% calibration
uncertainty added in quadrature with the likelihood-derived
errors.
We thank the Antarctic Support Associates staff at the South Pole
for making a successful season possible, and
Ted Griffith, Bob Pernic, and Bill Vinje
for valuable assistance there. This research was supported by the
James S. McDonnell Foundation,
PYI grant NSF AST 90-57089, and
the National Science Foundation
under a cooperative agreement with the Center for Astrophysical Research in
Antarctica (CARA), grant NSF OPP 89-20223. CARA is an NSF
Science and Technology Center. JR was supported by the McCormick
Fellowship at the University of Chicago.
\newpage
\begin{table}[h]
\centering
\begin{tabular}{c c c c} \hline \hline
Dataset& Field & Band Power & GACF \\
& & $T_0 \sqrt{\ell_e(\ell_e+1)C_\ell/(2\pi)}$ & $T_0 \sqrt{C_0}$ \\
& & $\ell_{e} = 93$ & $\theta_c = 1^\circ$ \\ \hline
PyI & A& $66^{+27}_{-11}$ & $90^{+36}_{-15}$ \\
PyII & A& $68^{+32}_{-11}$ & $91^{+42}_{-15}$ \\
(PyI\&PyII) & A& $69^{+27}_{-10}$ & $93^{+36}_{-13}$ \\
(PyI-PyII)/2.& A& $(3,19) $ & $(5,26)$ \\
PyII & B& $(16,59)$ & $(23,81)$ \\
(PyI\&PyII)&A\&B& $57^{+16}_{-8}$ & $79^{+23}_{-10}$ \\ \hline
\end{tabular}
\centering{
\parbox[htb]{5in}{
\caption{ Results of likelihood analyses.
\label{table_res}}
Values are in units of $\mu$K.
Detections are quoted at the maximum in the likelihood,
with 16\% of the integrated likelihood above the upper errorbar,
and 16\% of the integrated likelihood below the lower one.
For datasets with no significant detection, the value
which maximizes the likelihood is given, along with a
95\% upper limit from the integrated likelihood.
The 20\% calibration uncertainty is not included in these errors.
The signals from the dark channel and from the quadrature
phase are consistent with zero.
}
}
\end{table}
\newpage
|
\section{1. Introduction}
Canonical quantization of gauge systems has been a subject of much
discussion since the basic outline was first given by Dirac [5].
This formalism has been especially popular in the gravitational
physics community as, for Einstein's general theory of relativity on a
spatially compact universe, the Hamiltonian consists {\it only}
of constraints. In addition, the nature of the gauge transformations
associated with gravity
make gauge fixing techniques extremely difficult to apply and
perturbative nonrenormalizability has frustrated attempts at covariant
path integral quantization. Thus, Dirac style canonical quantization
remains at the forefront of quantum gravity research [2,10,11].
Despite this interest, certain basic issues
have remained unresolved for the general case. Recall that
the essential idea of Dirac's
approach is to turn the classical constraints $C_i$
into linear operators $\widehat{C}_i$ and to consider `physical
states $\psi_{phys}$' that are annihilated by the constraints; i.e., such
that $\widehat{C}_i \psi_{phys} = 0$.
However, the questions
of on which linear space the
constraints should act and of just how an inner product is to be
imposed on the solutions to define a Hilbert space do not yet have
widely accepted answers.
Recently, a resolution to these issues has been proposed.
In fact, what is
essentially the same resolution has been
independently
suggested twice under the names of the `Rieffel induction procedure'
[12] and the `refined algebraic quantization scheme' [1].
This method has been successfully used to quantize linearized
gravity on symmetric backgrounds [8,9], minisuperspace models for
gravity [16,17], and the free Maxwell field [13].
As might be expected, these methods proceed by introducing
additional structures beyond what is present in the original
Dirac approach. These techniques and additional structures
will be explored further here in the particular context of
systems with a single constraint.
We begin with a review of the Rieffel/refined algebraic procedure in
section 2. Here, we use the language and notation of [1]
as it is more closely related to that of Dirac [5] and therefore more
familiar. We will also refer to the scheme as the `refined
algebraic proposal' in the text.
Sections 3 and 4 contain the main results of this paper. In section 3
we show how, for typical systems with a single constraint,
a physical argument determines a unique implementation of
the Rieffel/refined algebraic scheme. In section 4
we discuss superselection laws on the physical Hilbert space and how their
existence may depend on the choice of auxiliary structures. We
give two examples. In the first, the use of an `incorrect' structure
leads to spurious superselection laws. This example also illustrates
the fact that the physical Hilbert space can depend
strongly on the choice of this structure.
In the second, the use
of a physically motivated auxiliary structure produces superselection
laws, but this time a similar feature exists in the classical theory.
Thus, for this second case we take the superselection laws to be
physically meaningful. Appendices A and B contain proofs of technical
results which are not of direct relevance to the main discussion but
which are mentioned in the text.
\section{2. The Refined Algebraic Approach}
In this section, we review the refined algebraic quantization scheme presented
in [1] (which is essentially equivalent to the Rieffel induction
procedure of [12]) for systems with gauge symmetries.
The starting point is a constrained classical system with phase space $\Gamma$
and, as usual, the nondegenerate symplectic form $\omega$ on
$\Gamma$ defines a Poisson Bracket on smooth functions $\Gamma \rightarrow
\Co$. The constraints $C_i$ are required to be first class; that is, the
Poisson bracket of two constraints is a sum of constraints
(possibly weighted by smooth phase space functions), as is the Poisson
bracket of any constraint with the Hamiltonian. As a result, the
constraint surface is preserved under time evolution.
The refined algebraic proposal quantizes this system in a series of steps
based on those of the original algebraic quantization program [2,3].
The first four steps below have nothing to do with constrained
systems but simply quantize the system obtained by ignoring the
constraints. They follow the unconstrained prescription of
[2,3] exactly but we repeat them
here for completeness and to fix our notation.
\vskip4pt plus2pt
\item {Step 1.} Select a subspace ${\bf S}$ of the vector space of
all smooth, complex-valued functions on $\Gamma$ subject to the
following conditions:
\vskip4pt plus2pt
\item \item a) $\bf S$ should be large enough so that any sufficiently
regular function on the phase space can be obtained as
(possibly a suitable limit of) a sum of products
of elements in $\bf S$.
\vskip4pt plus2pt
\item \item b) $\bf S$ should be closed under Poisson brackets,
i.e. for all functions $F, G$ in $\bf S$, their Poisson bracket
$\{F, G \}$ should also be an element of $\bf S$.
\vskip4pt plus2pt
\item \item c)
Finally, $\bf S$ should be closed under complex conjugation; i.e.
for all $F$ in $\bf S$, the complex conjugate $F^*$ should be
a function in $\bf S$.
\vskip4pt plus2pt
The idea is that each function in $\bf S$ is to be regarded
as an {\it elementary classical variable} which is to have
an {\it unambiguous} quantum analog.
\vskip4pt plus2pt
\item {Step 2.} Associate with each element $F$ in
$\bf S$ an abstract operator $\widehat F$. Construct the
free associative algebra generated by these {\it elementary quantum
operators}. Impose on it the canonical commutation relations,
$[\widehat F, \widehat G] = i \hbar \widehat{ \{ F, G \} }$, and, if necessary,
also a set of (anti-commutation) relations that captures the algebraic
identities satisfied by the elementary classical variables.
Denote the resulting algebra by ${\cal B}_{aux}$.
\vskip4pt plus2pt
\item {Step 3.} On this algebra, introduce an involution\footnote{${}^1$}{
Recall that an involution on ${\cal B}_{aux}$ is an anti-linear map $\star$ from
${\cal B}_{aux}$ to itself satisfying the following three conditions for all
$A$ and $ B$ in ${\cal B}_{aux}$: i) $(A + \lambda B)^\star = A^\star +
\lambda^* B^\star$, where $\lambda$ is any complex number; ii)
$(AB)^\star = B^\star A^\star$; and iii) $(A^\star)^\star = A$.
} operation
$\star$ by requiring that if two elementary classical variables $F$
and $G$ are related by $F^* = G$, then $\widehat F^\star = \widehat G$ in
${\cal B}_{aux}$. Denote the resulting $\star$-algebra by ${\cal
B}^{(\star)}_{aux}$.
\vskip4pt plus2pt
\item {Step 4.} Construct a linear $\star$-representation $R$ of
the abstract algebra $\bst_{aux}$ via linear operators on an
auxiliary Hilbert space $\H_{aux}$, i.e. such that
$$
R(\widehat A^\star) = R(\widehat{A})^\dagger
$$
for all $\widehat {A}$ in ${\cal B}^{(\star)}$, where $\dagger$ denotes Hermitian conjugation
with respect to the inner product in $\H_{aux}$.
\vskip4pt plus2pt
The remaining steps introduce the constraints and address the
questions raised in the introduction. That is, they first
use the space $\H_{aux}$ to provide a home for the constraints and for the linear
space on which they act, and then construct the physical Hilbert space
from the corresponding solutions.
\vskip4pt plus2pt
\item {Step 5a.} Represent the constraints $C_i$ as self-adjoint
operators $\widehat C_i$ (or, their exponentiated action, representing the
finite gauge transformations, as unitary operators $\widehat U_i$) on
$\H_{aux}$.
\vskip4pt plus2pt
We will now look for solutions
of the constraints in terms of {\it generalized} eigenvectors of $\widehat
C_i$ which will lie in the {\it topological dual} $\Phi'$ of some dense
subspace $\Phi \subset \H_{aux}$ (see also Ref. [6,7]). Since $\Phi$
and $\Phi'$ will be used to build the physical Hilbert space, we will
consider only physical operators that are well behaved with respect to
$\Phi$.
\vskip4pt plus2pt
\item {Step 5b.} Choose a suitable dense subspace $\Phi \subset
\H_{aux}$ which is left invariant by the constraints $\widehat{C_i}$ and let
$\bst_{phys}$ be the $\star$-algebra of operators on $\H_{aux}$ which commute
with the constraints $\widehat{C}_i$ and such that, for $A \in \bst_{phys}$, both
$A$ and $A^\dagger$ are defined on $\Phi$ and map $\Phi$ to itself.
\vskip4pt plus2pt
As an example in section 4 will illustrate, some physical input is in
general required to choose the space $\Phi$. Some factors governing this
choice are that
it must be sufficiently large so that $\bst_{phys}$ contains
`enough' physically interesting operators while it must also be
be sufficiently small that its topological dual $\Phi'$
contains enough physical states.
The main idea in the last few steps is that, while
the classical reality conditions
should determine the inner product, we should not need to explicitly
display a complete set of classical observables (i.e., functions which
Poisson commute with the constraints) to achieve this goal.
Instead, we use a
complete set of functions (${\bf S}$) on the {\it unconstrained}
phase space, noting that the reality properties of such functions will
determine the reality properties of the observables.
The reality conditions of operators in $\bst_{aux}$ are then implemented on the
auxiliary Hilbert space $\H_{aux}$. The physical Hilbert space
$\H_{phys}$ is to be constructed in such a way that any adjointness
relations involving only observables (i.e., $A = B^\dagger$, for $A,B$
observables) will in turn
descend from $\H_{aux}$ to $\H_{phys}$ (so that $A = B^\dagger$ on
$\H_{phys}$ as well). In this way, we will say that the
reality conditions are implemented on $\H_{phys}$.
We now wish to construct the physical Hilbert space $\H_{phys}$, which
will in general {\it not} be a subspace of $\H_{aux}$. The
key idea is to find an appropriate map $\eta: \Phi \rightarrow
\Phi'$ such that $\eta(\phi)$ is a solution of the constraints for all
$\phi \in \Phi$. (Note that the natural class of maps from $\Phi$ to
$\Phi'$ is {\it anti}-linear (e.g., the adjoint map)).
We proceed as follows.
\vskip4pt plus2pt
\item {Step 5c.} Find an anti-linear map $\eta$ from $\Phi$ to the
topological dual $\Phi'$ that satisfies:
\vskip4pt plus2pt
\item \item (i) For every $\phi_1 \in \Phi$, $\eta(\phi_1)$ is a solution
of the constraints; i.e.,
$$
0 = \bigl(\widehat{C}_i (\eta \phi_1)\bigr)[\phi_2] := (\eta \phi_1) [
\widehat{C}_i \phi_2]
$$
for any $\phi_2 \in \Phi$.
Here, the square brackets denote the natural action of $\Phi'$ on
$\Phi$.
\vskip4pt plus2pt
\item \item (ii) $\eta$ is real and positive in the sense that, for all
$\phi_1,\phi_2 \in \Phi$,
$$
({\eta \phi_1})[\phi_2] = ((\eta \phi_2)[\phi_1])^* {\ \ \ {\rm and}
\ \ \ }
$$
$$
(\eta \phi_1)[\phi_1] \geq 0 $$.
\vskip4pt plus2pt
\item \item (iii) $\eta$ commutes with the action of any $A \in \bst_{phys}$ in the
sense that
$$
(\eta \phi_1)[A \phi_2] = ((\eta A^\dagger \phi_1))[\phi_2]
$$
for all $\phi_1,\phi_2 \in \Phi$. The r.h.s. defines the so-called
dual action of $A$ on $\Phi'$ so that we may write this as
$\eta A \phi = A \eta \phi$.
\vskip4pt plus2pt
In analogy with [12] we call $\eta$ the {\it rigging map}.
(The appearance of the adjoint on the r.h.{s.} of the above equation
corresponds to the anti-linearity of $\eta$.)
\vskip4pt plus2pt
\item {Step 5d.} The vectors $\eta \phi$ span a space $\V_{phys}$ of
solutions of the constraints. We introduce an inner product on $\V_{phys}$
through
$$
\langle \eta \phi_1, \eta \phi_2 \rangle_{phys} = (\eta \phi_2) [ \phi_1]
$$
The requirement (iii) guarantees that this inner product is well
defined and that it is Hermitian and positive definite so that the
corresponding completion of $\V_{phys}$ is a
`physical' Hilbert space $\H_{phys}$.
(Note that the positions of $\phi_1$ and $\phi_2$ must be
opposite on the two sides of this definition due to the anti-linear nature
of $\eta$.)
\vskip4pt plus2pt
At this point, the reader may fear that this list of conditions on
$\eta$ will never be met in practice. That the new step 5 may
actually simplify the quantization program follows from the
observation of [8,9] (and [15,16] for the case when the
Poisson algebra of constraints is Abelian) that natural candidates
for such a map exist.
The last step is to represent
physical operators on $\V_{phys}$. This is straightforward because the
framework provided by step 5 guarantees that $\H_{phys}$ carries an (anti)
$\star$-representation (see below) of $\bst_{phys}$ as follows:
\vskip4pt plus2pt
\item {Step 6.} Operators in $ A \in \bst_{phys}$ have a natural action
(induced by duality) on $\Phi'$ that leaves $\V_{phys}$ invariant. Use this
fact to induce densely defined operators $A_{phys}$ on $\H_{phys}$ through
$$
A_{phys}\ (\eta \phi) = \eta ( A \phi).
$$
\vskip4pt plus2pt
This leads to an {\it anti}- $\star$-representation of $\bst_{phys}$ on
$\H_{phys}$ as
the map $A \mapsto A_{phys}$ from $\bst_{phys}$ to the operators on
$\H_{phys}$ is an anti-linear $\star$-homomorphism where $\star$ acts on the
operator $A_{phys}$ in the sense of conjugation of
quadratic forms on the dense domain $\Phi$
($\langle \phi, A^\star \psi \rangle \equiv \langle \psi,
A \phi \rangle^*$).
In this way, the reality
properties of the physical operators $\bst_{phys}$ on $\H_{aux}$ descend to the
physical Hilbert space.
In addition, consider any $C*$ algebra with unit ${\cal B}^{C*}$ which
is a subalgebra of $\bst_{phys}$.
Since the physical expectation value $\eta(A\phi)[\phi]$
defines a positive functional on ${\cal B}^{C*}$ (i.e.,
$\eta(A^\dagger A\phi)[\phi] \geq 0$), it follows that for $A
\in {\cal B}^{C*}$ we have
$$
\eta(A^\dagger A\phi)[\phi] \le ||A||^2
\eta(\phi)[\phi]
$$
so that $A_{phys}$ is a bounded operator on $\H_{phys}$ with norm
not larger than that of $A$ on $\H_{aux}$ ($||A||_{phys} \le ||A||$).
Thus, for such bounded operators, any relations of the form
$A = B^\dagger$ on $\H_{aux}$ also hold as the adjointness relations
$A_{phys} = B^\dagger_{phys}$ on $\H_{phys}$. From this it
follows that if $A$ is self-adjoint on $\H_{aux}$ and if a sufficiently
large class of bounded functionals of $A$ map $\Phi$ to itself, then $\bst_{phys}$
determines a (unique) self-adjoint extension of $A_{phys}$
on $\H_{phys}$.
Let us consider for a moment the case where there is only one
constraint. Note that when this constraint
has purely discrete spectrum, there is a natural choice for the map
$\eta$ as follows. Let $\Pi_0$ be the projection onto the eigenspace of the
constraint with eigenvalue zero. Then if we take $\Phi =
\H_{aux}$, the rigging map $\eta$ given by
$$
\eta |\psi \rangle = \langle \psi | \Pi_0
$$
fulfills all the requirements of step 5c. This case is simple and
easy to deal with, so that we shall focus on the complimentary
case (where the spectrum
is purely continuous) in the next section. Section 4 will describe
what happens when both continuous and discrete spectra are present.
\section{3. A Unique Prescription}
While the framework described in
section 2 sets the stage for quantizing constrained
systems, it does not provide the complete script. There are in fact
three inputs that need to be provided in order to proceed.
The first is the auxiliary space $\H_{aux}$ itself, but the dense subspace
$\Phi \subset \H_{aux}$ and the rigging map $\eta: \Phi \rightarrow \Phi'$
must also be given. As such, it is natural to ask to what extent
the above prescription is unique and to what extent it depends on
the choice of these inputs. In general, the answer is that the final
formulation depends a great deal on the inputs, as different
choices can even lead to physical Hilbert spaces of different dimensions!
This will be illustrated by an example in the next section.
Below, we confine ourselves to the case of a single
constraint $\widehat C$ of
the typical kind that arises in finite dimensional models.
The two main types of constrained systems are the `classic'
gauge systems in which the constraint is a vector field (whose orbits
are closed subsets of the configuration space) on some
configuration space and the `time reparametrization invariant
systems' in which the constraint is essentially the same as some
Hamiltonian of nonrelativistic quantum mechanics (but typically
with both positive and negative kinetic terms).
For such cases, physical reasoning will lead to a preferred choice of
the dense subspace $\Phi$ such that the rigging map
is then {\it unique} up to scale. As we consider constraints
with continuous spectrum, we shall assume that the configuration space
is noncompact. We argue as follows.
An important element of classical symplectic mechanics is that
the algebra of observables is taken to be the set of
{\it smooth} functions on
the phase space (as in step 1 of the refined algebraic program).
It is this definition, for example, that allows
us to talk about the (local) symplectomorphism `generated by an observable A.'
As such, the topology and differential structure
of the phase space play a key role and we would like to encode them
in our quantum formulation. Consider the
case where the classical phase space is $T^* \Rl^n$ and the auxiliary
Hilbert space used in the refined quantization program is $L^2(\Rl^n)$.
Recall that one characterization of the Schwarz space ${\cal S} \subset
\H_{aux}$ is as the set of all states $|\psi \rangle$ for which both $\langle x| \psi \rangle$
and $\langle p| \psi \rangle$ are smooth $L^2$ functions of $x$ and $p$,
where $\langle x|$ and $\langle p |$ are the usual position and momentum generalized
eigenstates. Thus, this set of states can be said to encode the
differential structure of the classical phase space and
is a natural choice for the subspace $\Phi$ of step 5b.
The algebra $\bst_{phys}$ of operators that
preserve this space contains all suitably
smooth and rapidly decreasing combinations of $x$ and $p$, in good
analogy with the classical algebra of observables. Thus, we take
$\Phi = {\cal S}$.
We will now show how a rigging map $\eta$ can be defined using
this choice and that this map is unique (given $\Phi =
{\cal S}$). Unfortunately, rigorous results are known to the author
only when certain additional assumptions are placed on the constraints
(which will be described below), but it is reasonable to conjecture that
similar results hold in the general case.
The result we need for our system is the following:
\vskip4pt plus2pt
\defin{Property A:}
{ There exists a set of generalized states $\langle c,k|$
for $c \in D_C$, $D_C$ an open subset of $\Rl$ containing $0$,
and $k \in D_K$, $D_K$ an open subset of $\Rl^{n-1}$, satisfying
$\langle c,k|\hat{C} = \langle c,k|c$ and $\langle c,k|c',k' \rangle
= \delta (c-c') \delta (k-k')$ and which are complete on the closed
subspace of ${\cal H}_{\rm aux}$ corresponding to the open spectral
interval $D_C$ of $\hat{C}$. The $\langle c,k|$ are
elements of the (algebraic) dual ${\cal S}^{\rm dual}$ to
${\cal S}$ and the map $F_k: c \mapsto \langle c,k|$ is continuous
with respect to the pointwise convergence topology on
${\cal S}^{\rm dual}$. Furthermore, the map $F: D_C \times {\cal S}
\rightarrow L^2(D_K,d^{n-1}k)$ given by $F: (c,|\psi\rangle)
\mapsto |\psi\rangle_c$ such that $\langle k|\psi\rangle_c
= \langle c,k| \psi \rangle$ is well-defined and smooth. }
\vskip4pt plus2pt
Such a result is easy to derive when the constraint is a vector field
with sufficiently regular orbits by simply introducing coordinates
in the space of orbits. For the case of a Hamiltonian constraint,
we will need to say something more about the form of the Hamiltonian.
Results are known for the following special cases:
\vskip4pt plus2pt
\item{1.} The massive free particle: Property A may be checked directly
using the momentum eigenstates.
\item{2.} The so-called separable semi-bound cases (see [17]): It follows
from the integral representation 5.14 of [17] that, when a scattering
operator exists for the `transverse' Hamiltonian $H_1$, there is
a complete set of orthogonal and appropriately normalized generalized
eigenstates $\langle c,k|$ satisfying Property A.
\item{3.} When the constraint is of the form $H = \sum_i p_i^2 + V(q) -E$
and $V\in L^1$: By extending Lemma IV.28 of [20] from
$C_0^\infty(\Rl^n)$ to ${\cal S}$, Property A reduces to the requirement that
$H$ have purely continuous spectrum.
\vskip4pt plus2pt
Unfortunately, the literature contains less helpful
results than one would like.
This is largely due to the fact that Hamiltonian constraints tend not
to have positive definite kinetic terms, while the literature is
primarily concerned with the Hamiltonians of particles moving on a
Riemannian space. Nevertheless, case 2 above contains nontrivial
cosmological models and
we suspect that Property A in fact holds in more general situations.
We will therefore assume that our system has Property A without
further justification.
Now, for $|\phi \rangle \in {\cal S}$, let $\phi(c,k)$
be the function $\langle c,k | \psi \rangle$.
Using Property A, we can construct the rigging map $\eta_0$ through
$$
(\eta_0 \phi_1)[\phi_2] = \int dc \ \delta(c) \ \int dk \
\phi_1^*(c, k) \phi_2(c,k)
$$
which clearly satisfies the criteria of step 5c. Note that the
action of the delta function is well defined since $\phi_1$ and
$\phi_2$ are continuous in $c$ by property A.
We will now see that this is the unique map (up to an overall scale)
that satisfies 5c. To do so, consider some generic rigging map $\eta$.
Since $\eta$ must commute
with the constraint, but has only solutions of the constraint in its
image, it is clear that $\eta$ must annihilate the the space $D \subset
{\cal S}$ of states which are in the domain of
${\widehat C}^{-1}$ and which are mapped into ${\cal S}$ by
${\widehat C}^{-1}$. This is the space of smooth $\phi(c,k)$ for which
$c^{-1} \phi(c,k)$ is also smooth. Since any smmoth function that vanishes
at $c=0$ at zero must vanish at least as fast as $c$, this is
in fact the space of all $|\phi \rangle
\in {\cal S}$ for which $\phi(0,k) = 0$. It follows that the kernel of
$\eta$ includes the kernel of $\eta_0$.
Let us now consider two states $|\phi_1\rangle, |\phi_2 \rangle \in {\cal S}$
which are {\it not} annihilated by $\eta_0$; that is, for which
$\phi_1 (0,k)$ and $\phi_2(0,k)$ are nonzero on a positive measure
subset of $D_K$. Then by
continuity there is some $\epsilon$ such that
$$ \int dk \ |\phi_i (c,k)|^2 > 0 \ \ \ \ (i = 1,2)$$
for all $|c| < \epsilon$ and such that $[-\epsilon, \epsilon]
\subset D_C$. We now define $\Pi_{[-\epsilon, \epsilon]}$
to be the projection onto the spectral interval $[-\epsilon, \epsilon]$
of the constraint $\widehat C$ and consider the state
$$
|\psi_1 \rangle = \Pi_{[-\epsilon, \epsilon]} |\phi_1 \rangle.
$$
Note that $|\psi_1 \rangle$ and $|\phi_1\rangle $ map to the same
element of $\Phi'$ under both $\eta$ and $\eta_0$.
We also define a state
$|\psi_2 \rangle$ by the equation
$$
\psi_2 (c,k) = \sqrt{ { {\int dk' |\phi_1(c,k')|^2} \over {\int dk'
|\phi_2(c,k')|^2} }} \phi_2 (c,k)
$$
for $ |c| \leq \epsilon$ and $\phi_2(c,k) = 0$ for $|c| > \epsilon$.
While $|\phi_2 \rangle$ and $|\psi_2 \rangle$ map to different elements of
$\Phi'$, they map to the same ray in ${\cal H}_{\rm phys}$ under $\eta$
and to the same ray in ${\cal H}_{{\rm phys},0}$ under $\eta_0$.
Note that $\eta_0 |\psi_1 \rangle$ and $\eta_0 |\psi_2\rangle$
have the same norm in ${\cal H}_{{\rm phys},0}$, but are otherwise
arbitrary elements of ${\cal H}_{{\rm phys},0}$.
We will now show that the conditions of step 5c guarantee that
$\eta|\psi_1 \rangle$ and $\eta |\psi_2 \rangle$ have the same physical
norm no matter how $\eta$ is defined. To proceed,
consider the family $U(\theta)$ of unitary operators
that generate rotations in the two dimensional subspace of $\H_{aux}$
spanned by $|\psi_1 \rangle$ and $|\psi_2 \rangle$ and note that, for
fixed $c$, the functions $\psi_i(c,k)$ define elements
$|\psi_{i,c}\rangle$ of the `transverse' Hilbert space $\h_c \sim L^2(D_K,dk)$.
Such $U(\theta)$
are in fact diagonal in $c$; that is, they satisfy
$$
\langle c,k |U(\theta) |\phi \rangle = \langle k | U_c(\theta) | \phi_c \rangle_c
$$
where the subscripts $c$ on the r.h.s. indicate that the matrix
element is taken in the transverse Hilbert space $\h_c$.
Here, $U_c(\theta)$ is just the unitary operator on $\h_c$
that rotates the subspace
spanned by $|\psi_{1,c} \rangle$ and $|\psi_{2,c}\rangle$ and $\langle k |$
is the ket for which $\langle k |\phi_c \rangle_c = \phi(c,k)$. As a result,
$U(\theta)$ commutes with the constraint $\widehat C$ and, since it preserves
the subspace $\Phi$, must belong to the algebra $\bst_{phys}$ of observables.
However, this means that it must commute with $\eta$ and define a
unitary operator on the corresponding physical Hilbert space.
It follows that whenever $(\eta_0 \psi_1)[\psi_1] = (\eta_0 \psi_2)[\psi_2]$,
we must also have
$(\eta \psi_1)[\psi_1] = (\eta \psi_2)[\psi_2]$. Since
$\eta$ provides a positive semidefinite inner product, the functional
$\phi \mapsto (\eta \phi)[\phi]$
in fact defines $\eta$ completely and $\eta$ must be just
$\eta_0$ up to some overall positive scale factor.
\section{4. Superselection Laws}
In contrast with the previous section, the case considered in
[1] did not result in a unique rigging map. Instead, a
large family of maps was found, associated with the existence of certain
`superselection rules.' It seems a
reasonable conjecture that, for a
given choice of subspace $\Phi$, the non-uniqueness
of the rigging map is always exactly determined by the
superselection rules. While we shall not prove this here, the
discussion
below provides supporting evidence. Appendix A shows that this is
true for the particular case studied in [1].
Interestingly, the very existence of superselection rules can
depend on the choice of the dense subspace $\Phi$ of step 5b.
This emphasizes the importance of choosing $\Phi$ based on physical
motivations. Below,
we provide two examples of cases where a
superselection
laws arises: one (in 4.1) in which it seems to come from the `wrong' choice
of $\Phi$, and one (in 4.2) in which its existence reflects
a feature of the classical physics.
{\bf 4.1} {\it The Dependence on $\Phi$.}
For our first example, we will rework the case of section 2 using a
different choice of $\Phi$. Property A allows us to introduce a
notion of {\it continuous} states as follows:
\vskip4pt plus2pt
\defin{Definition} {A state $|\phi \rangle \in \H_{aux}$ is said to be
continuous on $\Sigma \subset D_C$ if $\phi(c,k)$ is continuous in
$c$ for each fixed $k$ at every $c \in \Sigma$.}
\vskip4pt plus2pt
We will construct $\Phi$ in the following (complicated!) way. Choose
some interval \break $[-a,a] \subset D_C$. Now, consider the
subintervals $I_n^- = (-{a \over {2^n}}, - {a \over {2^{n+1}}})$
and $I_n^+ = ({a \over {2^{n+1}}},{a \over {2^n}})$
for $n \geq 0$.
Let $R_E$ be the union of the $I^\pm_n$ for even $n$ and $R_O$
be the union for odd $n$. In addition, consider a family
of projections $\Pi_c$ on $\h_c$ for which the matrix elements
$\langle k| \Pi_c | k' \rangle_c$ are independent of $c$ and let ${\cal N}_c$
be the subspace of $\h_c$ annihilated by $\Pi_c$.
We now let $\Phi$ be the dense subspace of $\h_{aux}$ containing all
states $|\psi \rangle$ such that
\vskip4pt plus2pt
\item A) $|\psi \rangle$ is continuous on $R_E$ and $\lim_{c \rightarrow 0 \
{\rm in} \ R_E}$ exists in $\Co$.
\item B) $|\psi \rangle$ is continuous on $R_O$ and $\lim_{c \rightarrow 0 \ {\rm
in} \ R_O}$ exists in $\Co$.
\item C) $\Pi_c |\psi_c\rangle = 0$ at the midpoint of $I^\pm_n$ for each
odd n.
\vskip4pt plus2pt
The limit in A (B) is taken by considering only
sequences in $R_E$ ($R_O$).
Note that since elements of $\Phi$ are only required to be continuous
separately on the sets $R_E$ and $R_O$, there are now ${\it two}$
natural choices for the rigging map, $\eta_E$ and $\eta_O$:
$$ (\eta_E \phi)[\psi] = \lim_{c \rightarrow 0 \ {\rm in} \ R_E}
\int dk \ \phi^*(c,k) \psi(c,k)$$
$$ (\eta_E \phi)[\psi] = \lim_{c \rightarrow 0 \ {\rm in} \ R_O}
\int dk \ \phi^*(c,k) \psi(c,k)$$
\vskip4pt plus2pt
\remar{Remark\ {1.}\ } {
Note that $\eta_E$ leads to the usual
physical Hilbert space $L^2(D_K,dk)$, whereas $\eta_O$ leads
to a smaller physical space isomorphic to ${\cal N}_c$.}
\vskip4pt plus2pt
\remar{Remark\ {2.}\ } {
For fans of group
averaging, we mention that the group averaging procedure [1,8,9,12]
does not converge on $\Phi$ (see Appendix B).}
\vskip4pt plus2pt
The existence of these two maps is associated with the following
superselection rule. Let $\Phi_E \subset \Phi$ contain
those states of $|\psi \rangle$ for which $\psi(c,k) = 0$ when
$c \in R_O$ and let $\Phi_O \subset \Phi$ contain those
for which $\psi(c,k) = 0$ when $c \in R_E$. Then, for any
$A \in \bst_{phys}$, because $[A,C] = 0$, we have
$\langle \phi_E | A | \phi_O \rangle = 0$ for any $|\phi_E \rangle \in \Phi_E$,
$|\phi_O \rangle \in \Phi_O$. Such superselection rules then
descend to the physical level; that is, to the
action of the physical operators on the physical
Hilbert space.
For a general constraint (such as, say, $p_x = 0$, generating
translation gauge invariance), there is no reason to expect
superselection rules. Also, the Hilbert space that results from
$\eta_O$ seems unreasonably small. Thus, we must regard these features
as artifacts of using the `wrong' choice of $\Phi$. In contrast, the
physically motivated choice of section 3 produced perfectly satisfactory
results.
{\bf 4.2} {\it Physical superselection laws}
We now turn an example the superselection
rule captures a feature of the corresponding classical
system, and thus appears physically meaningful.
For this case, we consider systems which differ
slightly from those considered so far. We now ask only that
our system satisfy `Property B:'
\vskip4pt plus2pt
\defin{Property B:}{ The Hilbert space $\H_{aux}$ can be written
as a direct sum $\H_{aux} = \h_{disc} \oplus \h_{cont}$
where $\h_{disc}$ is (densely) spanned by {\it normalizable} eigenstates
of $\widehat C$ and such that when the system is
restricted to $\h_{cont}$, it
satisfies Property A.}
\vskip4pt plus2pt
Now, let $\Phi = \h_{disc} \oplus \Pi_{cont} {\cal S}$ where
$\Pi_{cont}$ is the projection to ${\cal H}_{cont}$. Again, there are
two natural choices of rigging map. First is $\eta_{disc}$,
$$ \eta_{disc} |\psi \rangle = \langle \psi | \Pi_0$$ where $\Pi_0$ is
the projection onto the (normalizable) eigenstates of $\widehat C$ with
eigenvalue zero. Second is $\eta_{cont}$, defined to annihilate
$\h_{disc}$ but otherwise just as in section 3. Any combination
$a \eta_{disc} + b \eta_{cont}$ for $a,b > 0$ also defines a rigging
map that satisfies the requirements of step 5c.
Again, there is an associated superselection law between
$\h_{disc}$ and $\Pi_{cont} {\cal S}$. To see this,
note that since $A \in \bst_{phys}$ has an adjoint $A^\dagger \in \bst_{phys}$,
we need only show that, for all $A \in \bst_{phys}$, $A$ maps
$\h_{disc}$ into $\h_{disc}$
and we will be done. However, since $[A, \widehat C] =0$
and the domain of
$A$ contains $\Phi \supset \h_{disc}$, $A$
must map every normalizable eigenvector of $\widehat C$ to a normalizable
eigenvector of $\widehat C$ (with the same eigenvalue). Thus, each $A \in \bst_{phys}$
preserves $\h_{disc}$, providing us with a superselection rule.
Again, this descends to a superselection rule for the physical
operators on the physical Hilbert space.
However, this time the corresponding {\it classical} system has
a similar feature\footnote{${}^2$}{The argument given below is
an improved version of the one given in Appendix A of [1].}. To see this,
recall that when an operator $\widehat{A}$ is
associated with a function A on the classical phase space,
the discrete eigenvalues of the operator $\widehat{A}$ are
associated with parts of the phase space in which the orbits of
the Hamiltonian vector field of the function A are contained in
compact regions, while the continuous eigenvalues are associated with
parts of the phase space where these orbits are not contained in
compact regions.
Suppose then that we have a single classical constraint
$C$. For concreteness, we assume that the phase space $\Gamma$
is a finite dimensional manifold.
Let $\Gamma_{disc}$ be the union of the collection of all orbits $O$
generated by this constraint such that there exists a compact $K_O \in \Gamma$
containing $O$. We may think of $\Gamma_{disc}$ as the classical
analogue of the space $\h_{disc}$ of
discrete eigenvectors of $\widehat{C}$. Let $\Gamma_{cont}$
be the rest of the phase space $\Gamma$. Now, consider some
function $A$ on the phase space such that $A$ Poisson commutes with $C$.
The exponentiated action of the Hamiltonian vector field defined
by $A$ is a (local) homeomorphism that
maps orbits of $C$ onto orbits of $C$. Since, for an orbit
$O \subset \Gamma_{disc}$, every neighborhood $U \subset
\Gamma$ of $O$ contains some
compact set $K_U$ which contains $O$, we therefore conclude that
this exponentiated map cannot take an orbit $O_{disc} \in \Gamma_{disc}$ to an
orbit $O_{cont} \in \Gamma_{cont}$ and vice versa. Thus, we find that
(in the terminology of [14]) $\Gamma_{disc}$ and $\Gamma_{cont}$ contain
disjoint sets of symplectic leaves of $\Gamma$ and we have a
classical superselection law between the corresponding two parts of the
reduced phase space.
This seems to be the direct classical analogue of the quantum
superselection rules discussed above; in fact, it is even stronger.
All that is really required in the above argument is that the Poisson
bracket of $A$ and $C$ vanish on the constraint surface. Thus,
this superselection rule holds even for the so-called `weak
observables.'
It seems then that we must be careful. When the space $\Phi$
is chosen to reflect the smooth structure of the phase space, we
have found physically meaningful superselection rules, a reasonable
physical Hilbert space, and a (sufficiently) unique rigging map.
However, when this is not the rationale for choosing $\Phi$, spurious
results may occur. In the case of the diffeomorphism invariant
states of [1], the corresponding $\Phi$ was chosen to reflect
this structure as it is the appropriate domain of definition for the
operators that were assumed to function as coordinates and momenta.
Thus, within the framework of the auxiliary space of [1] and
modulo questions concerning the Hamiltonian constraint (which was
intentionally ignored), we expect that the superselection
rules of [1] should be taken seriously.
\section{Acknowledgments}
The author would like to thank Abhay Ashtekar, Petr Hajicek,
Atsushi Higuchi,
Nicholas Landsman, Jurek Lewandowski,
Jos\`e Mour\~ao, and Thomas Thiemann for many useful discussions.
Special thanks are due to Chris Isham and Karel Kucha\v{r}
for repeatedly asking about the uniqueness of the physical Hilbert
space given by the refined algebraic approach, to Chris
Fewster for especially clarifying discussions and for help in locating
reference [19], and to Carlo Rovelli for discussions on the
significance of the superselection rules. Finally, many thanks to
Domenico Giulini for pointing out an error in a previous version of
the paper.
\section {Appendix A. Uniqueness of the construction of Connection
Representation Diffeomorphism Invariant States}
In this appendix, we give a short proof that the rigging maps used
in [1] to solve the diffeomorphism
constraint completely exhaust the set of possible such
maps given the choice of auxiliary space, the definitions of the quantum
constraints, and the dense subspace $\Phi$ chosen in [1].
For a full definition of the terms and notation used below, see
[1].
Recall from that the auxiliary Hilbert space of [1]
is spanned by a set of orthonormal `spin network states.' We
shall denote these states by $|\Gamma_{\alpha, k} \rangle$,
where $\alpha$ is a (piecewise
analytic) graph embedded in a given analytic three manifold and $k$
is an index that takes some finite set of values (this set depends
on the graph $\alpha$). In addition, (analytic)
diffeomorphisms ${\cal D}$ act on these states by moving the graph
$\alpha$ in the obvious way and permuting the values of the index $k$
allowed by $\alpha$.
The dense subspace $\Phi$ of step 5b
is the space of so-called smooth cylindrical
functions. This space contains all finite linear combinations of the
spin network states $\Gamma_{\alpha, k}$ and, for our purposes,
may in fact be identified with this slightly smaller space.
Following [2], we shall consider only `type I graphs' (see [2]).
As in [1], it is convenient to introduce the subspaces
$\h^{[\tilde \beta]}$ spanned by spin networks $|\Gamma_{\alpha,k}\rangle$
associated with
graphs $\alpha$ that can be moved by a diffeomorphism to some
graph $\beta$ for which $\tilde \beta$ is the `maximal analytic
extension.' These subspaces are superselected by the algebra
$\bst_{phys}$ and, on each subspace, there is a corresponding map
$\eta^{[\tilde \beta]}$ defined by:
$$
\eta^{[\tilde{\beta}]} |f\rangle =
\bigl( \sum_{{\cal D}_1 \in S(\tilde \beta)}
\sum_{[{\cal D}_2] \in GS(\tilde \beta)} {\cal D}_1 {\cal D}_2 |f \rangle
\bigr)^\dagger
$$
where we still need to introduce the set $S (\tilde \beta)$
and the quotient space
$GS(\tilde \beta)$. $S(\tilde \beta)$ is chosen to be any set (and the above
map does not depend on this choice) of diffeomorphisms ${\cal D}_{
\tilde \alpha}$, one for each maximally
extended analytic graph $\tilde \alpha$
diffeomorphic to $\tilde \beta$, such that ${\cal D}_{\tilde \alpha}$
moves the extended graph $\tilde \beta$ onto the extended graph $\tilde
\alpha$. On the other hand, $GS(\tilde \beta)$ (the `graph
symmetry group' of $\tilde \beta$) is the quotient $Iso(\tilde \beta)
/ TA(\tilde \beta)$ where the `isotropy
group' $Iso(\tilde \beta)$ is the group of diffeomorphisms
which map $\tilde \beta$ onto $\tilde \beta$ and the `trivial
action group' $TA(\tilde \beta)$ is the group of diffeomorphisms
that map every edge $e$ in $\tilde \beta$ onto itself.
In the formula above, $[{\cal D}_1]$ denotes the equivalence class of
${\cal D}_2$ in $GS(\tilde \beta)$.
Any linear combination $\sum_{i \in I} a_i \eta^{[\tilde \beta_i]}$ with
positive coefficients $a_i$ satisfies the requirements of step
5c. (Note that this sum always converges no matter how big the
coefficients $a_i$ or the index set $I$.)
We would now like to show that such sums exhaust the set of
all rigging maps.
We will follow the same basic strategy as in the uniqueness proof of
section 3. That is, we now consider a generic map $\eta$ satisfying
5c and show that if $\eta^{[\tilde
\beta]} |\phi \rangle = 0$ for all $\tilde \beta$, then $\eta |\phi\rangle = 0$
as well.
Suppose then that
such that $\eta^{[\tilde \beta]} |\phi_0\rangle = 0$ for all
$\tilde \beta$.
Since $|\phi_0 \rangle \in \Phi$, it can be written as a sum of
spin network states. It will be particularly convenient to
write it in the form:
$$
|\phi_0 \rangle = \sum_{i} \sum_j c_{ij} {\cal D}_j |\Gamma_{
i} \rangle
$$
where $c_{ij} \in \Co$, ${\cal D}_j \in {\rm Diff}^\omega$, and
$\{ |\Gamma_{i}\rangle \}$ is some set of spin network states,
carefully chosen so that no analytic diffeomorphism maps one
spin network state in this set onto another.
Now, it is easily checked that $|\phi_0 \rangle$ is annihilated by
the above maps exactly when $\sum_j c_{ij} = 0$ for each $i$.
However, any rigging map that commutes with diffeomorphisms and whose
image contains only diffeomorphism invariant states must also
annihilate states with $\sum_j c_{ij} = 0$. Thus, $\eta |\phi_0\rangle = 0$.
Now consider some spin network state $|\Gamma_0\rangle :=
|\Gamma_{\alpha, k}\rangle$ such that
$\eta |\Gamma_0\rangle$ is nonzero (so that $\eta^{[\tilde \alpha]} |\Gamma_0\rangle$ is nonzero
as well) and choose any other state
$|\Gamma_1\rangle \in \h^{[\tilde \alpha]} \cap \Phi$. We want to construct an
operator $A$ in $\bst_{phys}$ that has nonzero matrix elements between
$|\Gamma_0\rangle$ and $|\Gamma_1\rangle$. This can be done by applying just
the kind of `group averaging' that was used in the construction of
$\eta^{[\tilde \beta]}$:
$$
A:= \sum_{{\cal D}_1 \in S(\tilde \alpha)} \sum_{[{\cal D_2}] \in
GS(\tilde \alpha)}
{\cal D}_1 {\cal D}_2 |\Gamma_1 \rangle \langle \Gamma_0|
{\cal D}_2^{-1} {\cal D}_1^{-1}.
$$
This operator is diffeomorphism invariant and finite on $\Phi$ for
exactly the same reasons as the map $\eta^{[\tilde \alpha]}$ (and similarly for $A^{\dagger}$).
As a result, it is an element of $\bst_{phys}$.
Note that $A|\Gamma_0\rangle$ is a sum of spin networks that differ from
$|\Gamma_1\rangle$ only by a diffeomorphism. Thus, $A|\Gamma_0\rangle$
maps under $\eta$ to a diffeomorphism invariant state that is proportional
to $\eta |\Gamma_1\rangle$. However, the number of terms in this sum is
just the physical norm of the state $|\Gamma_0\rangle$ as defined through
the map $\eta^{[\tilde \alpha]}$ (and similarly for $A^{\dagger}|\Gamma_1\rangle$).
Let us therefore set
$N_0 = (\eta^{[\tilde \alpha]} \Gamma_0)[\Gamma_0]$ and $N_1 = (\eta^{[\tilde \alpha]} \Gamma_1)[\Gamma_1]$
so that $\eta A^{\dagger} A |\Gamma_0\rangle
= N_0 \eta A^{\dagger} |\Gamma_1\rangle = N_0 N_1 \eta |\Gamma_0\rangle$.
Applying this distribution to $|\Gamma_0 \rangle$ we have:
$$
{{ (\eta \Gamma_1)[\Gamma_1]} \over {N_1} } = {{(\eta \Gamma_0)[\Gamma_0]}
\over {N_0}}.$$ As before, this guarantees that when acting on
the subspace $\h^{[\tilde \alpha]} \cap \Phi$, $\eta$ acts just
like $\eta^{[\tilde \alpha]}$ up to an overall positive scale factor. Since the
domains of the $\eta^{[\tilde \alpha]}$'s are orthogonal,
it follows that $\eta$
may in fact be expressed as a sum
of the $\eta^{[\tilde \alpha]}$ weighted by positive coefficients.
\section{Appendix B. Convergence of the group averaging procedure}
In this appendix we show that the integral that defines the
group averaged inner product does not (absolutely) converge on
the entire space $\Phi$ given in the second example of section 4.
Recall that the group averaging proposal [1,8,9,12]
is to introduce the physical inner product
$$
\langle \phi, \psi \rangle_{phys} = \int dt \langle \phi, e^{i t \widehat{C}}
\psi \rangle_{phys} $$
for $\phi, \psi$ in $\Phi$. If this integrand is in fact $L^1$, then
we may write this as
$$
\lim_{T \rightarrow \infty} \int_{-T}^T \langle \phi, e^{i t \widehat{C}}
\psi \rangle = \lim_{T \rightarrow \infty}
\langle \phi, { {\sin (T\widehat{C} )} \over {\widehat{C}}} \psi \rangle
$$
$$
{\hskip 3 cm} = \lim_{T \rightarrow \infty}
\int_{\Lambda} d \lambda \langle \phi (\lambda),
{ {\sin (T\lambda)} \over {\lambda}} \psi (\lambda) \rangle_\lambda.
$$
However, we will now show that this limit fails to exist for general
$\phi, \psi \in \Phi$.
For convenience, we assume that $D_C = \Rl$. Furthermore, we
will take $a= 1$ and introduce the intervals $J_n^-
= (-2^{n+1}, -2^n)$, $J_n^+ = (2^n,2^{n+1})$. Finally, let
$R'_E = R_E \cup (\cup_{\pm, {\rm even} \ n} J_n^\pm)$,
$R'_O = R_O \cup (\cup_{\pm, {\rm odd} \ n} J_n^\pm)$,
and
let
$|\psi \rangle \in \Phi$ be any state such that $\psi (c,k)$
vanishes for $c$ in $R'_O$.
The important
property of $R'_E$ is that this set is preserved when the
real line is scaled by a factor of $2^k$. As such, given any
function $\psi(\lambda)$ which is continuous on $R'_E$,
the limit
$$
\lim_{k \rightarrow \infty}
\int_{R'_E} { {\psi(\lambda) \sin (2^k T \lambda)} \over
{\lambda}}
$$
for large $k$ is just
$$
\psi(0) \int_{R'_E} { { \sin (2^k T \lambda)} \over
{\lambda}} \equiv \psi(0) I(T)
$$
which is independent of $k$.
It follows that the limit
exists for large $T$ if and only if $I(T)$ is constant.
However, we will now show that $I(T)$ is not constant. Note
that its derivative is
$$dI/dT = \int_{R'_E} \cos(T \lambda) d \lambda$$
and suppose that $T=\pi$. Then,
$\int_{J^\pm_n} \cos(\pi \lambda) d\lambda = 0$, but
we have $\int_{I^\pm_n} \cos(\pi \lambda) >0$ so that $I(\pi) > 0$.
As a result, the group averaging norm does not
(absolutely) converge for any nontrivial $|\psi \rangle \in \Phi$
that vanishes on $R'_O$.
\references{}{
\item{[1]} A. \spa{Ashtekar}, J. \spa{Lewandowski}, D. \spa{Marolf},
J. Mour\~ao, and T. \spa{Thiemann}, {\it Quantization of
diffeomorphism invariant theories of connections with local
degrees of freedom}\/,
J.~Math. Phys. {\bf 36} 6456 (1995); gr-qc/9504018
\item{[2]} A. \spa{Ashtekar}, {\it Non-Perturbative Canonical Gravity},
Lectures notes prepared in collaboration with R.S. Tate, World Scientific,
Singapore, 1991.
\item{[3]} A. \spa{Ashtekar} and R. S. \spa{Tate}, {\it
An algebraic extension of Dirac quantization: Examples},
J. Math. Phys. 35 (1994) 6434.
\item{[4]} B.~\spa{DeWitt}, {\it Quantum Theory of Gravity.I. The
Canonical Theory} Phys. Rev. 160 (1967), 1113
\item{[5]} P.A.M. \spa{Dirac} {\it Lectures on Quantum Mechanics}\/,
Belfer Graduate School of Science, Yeshiva University,
New York, 1964
\item{[6]} I.M. \spa{Gel'fand}, N.Ya. \spa{Vilenkin}, {\it
Generalized Functions: vol. 4, Applications of Harmonic Analysis},
Academic Press, New York, London, 1964.
\item{[7]} P. \spa{Hajicek}, {\it Quantization of Systems with
Constraints}
in {\it Canonical Gravity: from
classical to quantum} ed. by
J. \spa{Ehlers}, H. \spa{Friedrich}, Lecture notes in Physics,
Springer-Verlag, Berlin, New York, 1994.
\item{[8]} A. \spa{Higuchi}, {\it Quantum linearization instabilities
of de Sitter spacetime: II}, Class. Quant. Grav. 8 (1991) 1983.
\item{[9]} A. \spa{Higuchi}, {\it
Linearized quantum gravity in flat space with
toroidal topology},
Class. Quant. Grav. {\bf 8} (1991) 2023.
\item{[10]} C.~\spa{Isham}, {\it Canonical Gravity and the Problem of
Time}, Imperial College, preprint TP/91-92/25, gr-qc/9210011 (1992).
\item{[11]} K. Kucha\v{r}, {\it Time and Interpretations of
Quantum Gravity} in {\it Proceedings of the 4th Canadian
Conference on General Relativity and Relativistic Astrophysics},
ed. G. \spa{Kunstatter} et. al.
World Scientific, New Jersey 1992.
\item{[12]} N. \spa{Landsman}, {\it Rieffel induction as
generalized quantum Marsden-Weinstein reduction}\/,
J. Geom. Phys. 15 (1995) 285-319; hep-th/9305088.
\item{[13]} N. \spa{Landsman} and U. \spa{Wiedemann}, {\it Massless Particles,
Electromagnetism, and Rieffel Induction}\/,
Rev. Mod. Phys. {\bf 7} 923 (1995); hep-th/9411174.
\item{[14]} N.~\spa{Landsman}, {\it Classical and quantum representation
theory}, in
{\it Proceedings Seminar Mathematical Structures in Field Theory},
E. A. de Kerf and H.G.J. Pijls (eds) (CWI-syllabus, CWI, Amsterdam,
to appear 1995).
\item{[15]} D. \spa{Marolf}, {\it The spectral analysis inner product for
quantum gravity}, preprint gr-qc/9409036, to appear in the
Proceedings of the VIIth Marcel-Grossman Conference, R. Ruffini and
M. Keiser (eds) (World Scientific, Singapore, 1995); D. Marolf,
{\it Green's Bracket Algebras and their Quantization},
Ph.D. Dissertation, The University of Texas at Austin (1992).
\item{[16]} D. \spa{Marolf}, {\it Quantum observables and recollapsing
dynamics}, Class. Quant. Grav. 12 (1995) 1199,
gr-qc/9404053.
\item{[17]} D.~\spa{Marolf}, {\it
Observables and a Hilbert Space for
Bianchi IX}, Class. Quant. Grav. 12 (1995), 1441; gr-qc/9409049.
\item{[18]}
D. \spa{Marolf} {\it Almost Ideal Clocks in Quantum Cosmology: A Brief
Derivation of Time}, Class. Quant. Grav. {\bf 12} 2469 (1995);
gr-qc/9412016.
\item{[19]} B. \spa{Simon}, {\it Quantum Mechanics for Hamiltonians
defined as quadratic forms}, Princeton Univ. Press, Princeton, 1971,
p. 120.
}
\bye
|
\section{Introduction}
The purpose of this article is to describe the general analytic solution to
the functional equation
\begin{equation}
\phi_1(x+y)=
{ { \biggl|\matrix{\phi_2(x)&\phi_2(y)\cr\phi_3(x)&\phi_3(y)\cr}\biggr|}
\over
{ \biggl|\matrix{\phi_4(x)&\phi_4(y)\cr\phi_5(x)&\phi_5(y)\cr}\biggr|} }.
\label{functional}
\end{equation}
Although this equation appears to depend on five a priori unknown functions
we shall show that (\ref{functional}) is invariant under a
large group of symmetries ${\cal{G}}$ and that each orbit has a
solution of a particularly nice form, expressible in terms elliptic functions:
\begin{thm}
The general analytic solution to the
functional equation (\ref{functional})\ is, up to a ${\cal{G}}$ action given
by (\ref{symm}-\ref{symmsb}), of the form
$$\phi_1(x)= {\Phi(x;\nu_1)\over\Phi(x;\nu_2)},\quad
{\phi_2(x)\choose\phi_3(x)}={\Phi(x;\nu_1)\choose\Phi(x;\nu_1)\sp\prime}\quad
{ and}\quad
{\phi_4(x)\choose\phi_5(x)}={\Phi(x;\nu_2)\choose\Phi(x;\nu_2)\sp\prime}.
$$
Here
\begin{equation}
\Phi(x;\nu)\equiv {\sigma(\nu-x)\over {\sigma(\nu)\sigma(x)}}\,
e\sp{\zeta(\nu)x}
\label{soln}
\end{equation}
where $\sigma(x)=\sigma(x|\omega,\omega\sp\prime)$ and
$\zeta(x) ={\sigma(x)\sp\prime \over\sigma(x)}$
are the Weierstrass sigma and zeta functions.
\end{thm}
The group ${\cal G}$ of symmetries of (\ref{functional})\ will
be described further below. Our proof is constructive and indeed yields
more:
\begin{thm}
Let $x_0$ be a generic point for (\ref{functional}).
Then (for $k=1,2$) we have
\begin{eqnarray*}
\partial_y \ln
\left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(y+x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(y+x_0)\cr}
\right|
\Biggl|_{y=0}
&=&
\zeta(\nu_k)-\zeta(x)-\zeta(\nu_k-x)-\lambda_k
\\
&=&
-{1\over x}-\lambda_k +\sum_{l=0}F_l\, {x\sp{l+1}\over (l+1)!}
\end{eqnarray*}
and the Laurent expansion
determines the parameters $g_1$, $g_2$ (which are the same for both $k=1,2$)
characterising the elliptic functions of (\ref{soln}) by
$$
g_2={5\over3}\left({F_2+6 F_0\sp2 }\right),
\quad\quad
g_3= 6 F_0\sp3 -F_1\sp2 +{5\over3}F_0 F_2,
$$
and the parameters $\nu_k$ via $F_0=-\wp(\nu_k)$. Further, we have
$$
\phi_1(x+2 x_0)= \phi_1(2 x_0)\,
e\sp{(\lambda_2-\lambda_1) x}\,{\Phi(x;\nu_1)\over\Phi(x;\nu_2)}\,
$$
and
$$
\left( \matrix{\phi_{2k}(x+x_0)\cr \phi_{2k+1}(x+x_0)\cr} \right)
=
{e\sp{-\lambda_k x}\over f(x)}
\Biggl(
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\Biggr)
\Biggl( \matrix{1&0\cr \lambda_k&-1\cr}\Biggr)
\Biggl( \matrix{\Phi(x;\nu_k)\cr \Phi\sp\prime(x;\nu_k)\cr}\Biggr).
$$
Here the function
$$
f(x)=
{e\sp{-\lambda_k x}\over \Phi(x;\nu_k)}\,
{
\left|
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\right|
\over
\left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(x_0)\cr}
\right|
}
$$
is in fact the same for $k=1,2$.
\end{thm}
The term \lq generic\rq\ will be defined below and we will give
more expressions for the quantities appearing in the theorem.
One merit of writing (\ref{functional}) in this general form is that
several different functional equations may now be seen as
different points on a ${\cal G}$-orbit of
(\ref{functional}). Thus for example
\begin{eqnarray}
\phi_1(x+y)&=&\phi_1(x)\phi_1(y)\label{exponential}\\
\phi_1(x+y)&=&\phi_4(x)\phi_5(y)+\phi_4(y)\phi_5(x)\label{dexponential}\\
A(x+y)[ B(x)-B(y)]&=&A(x) A\sp\prime(y) -A(y)A\sp\prime(x).
\label{calfun}
\end{eqnarray}
are particular\footnote{
These correspond to
\begin{itemize}
\item[(a)]$\phi_2(x)=\phi_1(x)\phi_4(x)$ and $\phi_3(x)=\phi_1(x)\phi_5(x)$,
\item[(b)]$\phi_2(x)=\phi_4\sp2(x)$ and $\phi_3(x)=\phi_5\sp2(x)$,
\item[(c)]$\phi_1(x)=\phi_2(x)=A(x)$, $\phi_3(x)=A\sp\prime(x)$,
$\phi_4(x)=B(x)$ and $\phi_5(x)=1$.
\end{itemize}
} examples of (\ref{functional}).
The functional equation for the exponential (\ref{exponential})
corresponds to $\nu_1=\nu_2$ in our solution and the exponential comes
wholly from ${\cal G}$.
Particular cases of (\ref{dexponential}) have been studied in
\cite{BK} and we shall determine (see Lemma 4)
the general solution to (\ref{dexponential}) as an application of our work.
More interesting is equation (\ref{calfun}) which has been
studied by various authors
with assumptions of even/oddness on the functions appearing
\cite{Ca2, OPc, PS} or assumptions on the nature of $B$ \cite{Kr1}.
The general solution \cite{BCb, Bu1} $A(x)=\Phi(x;\nu)$ now
corresponds to to the limit $\nu_2\rightarrow0$ together with a
${\cal G}$ action. This will be illustrated later.
Finally, when
$\phi_1(x)=\alpha(x)$, $\phi_2(x)=\alpha(x)\tau(x)$, $\phi_4(x)=\tau(x)$,
$\phi_3(x)=\phi_2\sp\prime(x)$ and $\phi_5(x)=\phi_4\sp\prime(x)$
we obtain the functional equation
\begin{equation}
\alpha(x+y)=\alpha(x)\alpha(y)+\tau(x)\tau(y)\psi(x+y).
\label{calogero}
\end{equation}
The function $\psi(x)$ will be described in more detail in the sequel.
This equation was studied by Bruschi and Calogero \cite{BCb}
and will be used in our analysis.
Let us remark that both (\ref{dexponential}) and (\ref{calogero})
may be viewed as limiting cases of the functional equation
\begin{equation}
\Psi _1(x+y)=\Psi _2(x+y) \phi_2(x)\phi_3(y) +\Psi_3(x+y) \phi_4(x)\phi_5(y),
\label{eq:biggy}
\end{equation}
which \textit{a priori } depends on seven unknown
functions. Later we shall show how (\ref{functional}) may be used to solve
this.
It remains to place (\ref{functional}) in some form of context.
The last decade has seen a remarkable confluence of ideas from
completely integrable systems, geometry, field theory and functional
equations that is still being assimilated. To make some of these matters
concrete let us consider how such functional equations arise in the
context of integrable systems of particles on the line.
A pair of matrices ${L,M}$ such that ${\dot L}= [L,M]$ is known as a
Lax Pair; this is a zero curvature condition.
Starting with an ansatz for the matrices $L$ and $M$ one
seeks restrictions necessary to obtain equations of motion of some desired
form. These restrictions typically involve the study of functional equations.
The paradigm for this approach is the Calogero-Moser \cite{Ca}
system.
Beginning with the ansatz (for $n\times n$ matrices)
$$L_{jk}=p_j\delta_{jk}+ \, g\,(1-\delta_{jk}) A(q_j-q_k),\quad
M_{jk}=g\, [\delta_{jk}\sum_{l\ne j}B(q_j-q_l)-(1-\delta_{jk})C(q_j-q_k) ]
$$
one finds ${\dot L}= [L,M]$
yields the
equations of motion for the Hamiltonian system ($n\ge3$)
$$H={1\over 2}\sum_{j}p_j\sp2 +g\sp2\sum_{j<k}U(q_j-q_k)
\quad\quad\quad U(x)=A(x) A(-x) + {\rm constant}$$
provided $C(x)=-A\sp\prime(x)$,
and that $A(x)$ and $B(x)$ satisfy the functional equation (\ref{calfun}).
With this ansatz and assuming $B(x)$ even\footnote{
This assumption can in fact be removed \cite{BB}. }
Calogero \cite{Ca2} found
$A(x)$ to be given by (\ref{soln}). In this case the corresponding potential
is the Weierstrass $\wp$-function: $A(x)A(-x)=\wp(\nu)-\wp(x)$.
The functional equation (\ref{calogero}) is associated with a different
ansatz and yields the relativistic Calogero-Moser systems \cite{BCa, Ra}.
Similarly (\ref{functional}) arises from a more general ansatz \cite{BB}
associated with equations of motion of the form
$$
{\ddot q_j}=\sum_{k\ne j}(a+b \dot q_j) (a+b \dot q_k)
V_{jk}(q_j-q_k),
$$
which combines both relativistic ($b\ne0$) and nonrelativistic
($b=0$) systems together with potentials that can vary between particle pairs.
This unifies, for example, Calogero-Moser and Toda systems \cite{RS, Ra, Rb}.
The relativistic
examples yield the functional equation (\ref{functional}) while the
nonrelativistic situation involves the functional equation
\begin{equation}
\phi_6(x+y)=\phi_1(x+y)\big( \phi_4(x)-\phi_5(y)\big)
+{ \biggl|\matrix{\phi_2(x)&\phi_3(y)\cr\phi\sp\prime_2(x)&\phi\sp\prime_3(y)
\cr}\biggr|}.
\label{functional2}
\end{equation}
The general analytic solution to (\ref{functional2}) has yet to
be determined although particular solutions are known.
We remark that (\ref{functional}), and after suitably
symmetrizing (\ref{functional2}), are
particular cases of the functional equation
\begin{equation}
\sum_{i=0}\sp{N}
\phi_{3i}(x+y)
{ \biggl|\matrix{\phi_{3i+1}(x)&\phi_{3i+1}(y)\cr
\phi_{3i+2}(x)&\phi_{3i+2}(y)\cr}\biggr|}=0
\label{Bfun}
\end{equation}
with $N=1$ in the former case and $N=2$ in the latter.
When $\phi_{3i+2}=\phi_{3i+1}\sp\prime$
Buchstaber and Krichever have discussed (\ref{Bfun}) in connection with
functional equations satisfied by Baker-Akhiezer functions \cite{BKr}.
Lax pairs are but one way in which functional equations are associated
with integrable systems and we mention \cite{BFV, Ca3, BP, GT} for others.
There also appears a close connection between these functional
equations and the elliptic genera associated with the
string inspired Witten index \cite{HBJ, Kr}. Krichever for example
used the functional equation (\ref{calfun}) in his proof
of the \lq rigidity\rq\ property of elliptic genera \cite{Kr}
and it also appears when discussing rational and pole solutions
of the KP and KdV equations \cite{Kr1, AMM}.
We feel this connection between functional equations and
completely integrable systems is part of a broader and less well
understood aspect of the subject that deserves further attention.
An outline of the paper is as follows. First we will discuss the group of
symmetries of (\ref{functional}). These will be used in the proof of
theorem 1. Before turning to the proof we show in section 3
how the indicated solution
does indeed satisfy (\ref{functional}), using this as a vehicle to recall
some of the properties of elliptic functions that we will need throughout.
Section 4 is devoted to the proof of theorem 1 and
section 5 to that of theorem 2. Several applications of our theorems
including the general analytic solution to (\ref{dexponential})
and a discussion of (\ref{eq:biggy})
are then given in section 6. An appendix is given containing
various elliptic function formulae we shall make use of.
Various versions of Theorem 1 have appeared in unpublished preprints.
In \cite{Bu2} the form of $\phi_1(x)$ only was stated. In
\cite{BB} we introduced the $\cal G$ action to give Theorem 1 in
its present form. In improving the proof of this we obtained Theorem 2,
given here alongside the better proof of Theorem 1.
\section{The Group of Symmetries}
We next describe the group ${\cal G}$ of
invariances of (\ref{functional}).
Theorem $1$ gives a representative of each ${\cal G}$ orbit
on the solutions of (\ref{functional})\ with a particularly nice form.
First observe that a large group of symmetries ${\cal G}$ act on the solutions
of (\ref{functional}). The transformation
\begin{equation}
\Biggl( \phi_1(x), {\phi_2(x)\choose\phi_3(x)},{\phi_4(x)\choose\phi_5(x)}
\Biggr) \rightarrow
\Biggl( c\, e\sp{\lambda x} \phi_1(x), U {e\sp{-\lambda\sp\prime x}
\phi_2(x)\choose e\sp{-\lambda\sp\prime x}\phi_3(x)},
V {e\sp{\lambda\sp{\prime\prime}x} \phi_4(x)\choose
e\sp{\lambda\sp{\prime\prime}x} \phi_5(x)} \Biggr)
\label{symm}
\end{equation}
clearly preserves (\ref{functional})\ provided
\begin{equation}
\lambda+\lambda\sp\prime+\lambda\sp{\prime\prime}=0,\quad\quad
U,V\in GL_2,\quad\quad {\rm and}\quad\quad \det U=c\,\det V.
\label{constraints}
\end{equation}
Further, (\ref{functional}) is also preserved by
\begin{equation}
\Biggl( \phi_1(x), {\phi_2(x)\choose\phi_3(x)},{\phi_4(x)\choose\phi_5(x)}
\Biggr) \rightarrow\Biggl(
{1\over \phi_1(x)},{\phi_4(x)\choose\phi_5(x)},{\phi_2(x)\choose\phi_3(x)}
\Biggr)
\label{symms}
\end{equation}
and
\begin{equation}
\Biggl( \phi_1(x), {\phi_2(x)\choose\phi_3(x)},{\phi_4(x)\choose\phi_5(x)}
\Biggr) \rightarrow\Biggl(\phi_1(x),
f(x){\phi_2(x)\choose\phi_3(x)},f(x){\phi_4(x)\choose\phi_5(x)}\Biggr).
\label{symmsb}
\end{equation}
We will use these symmetries in our proof of theorem $1$ to find a
solution of (\ref{functional})\ on each ${\cal G}$ orbit
with a particularly nice form.
\section{Illustration of the Solution}
Before proceeding to the proof it is instructive
to see how the stated solution satisfies (\ref{functional}).
This will also allow us to introduce some elliptic function identities needed
throughout.
From the definition of the zeta function we have
\begin{equation}
\bigr(\ln \Phi(x;\nu)\bigl)\sp\prime = -\zeta(\nu-x)-\zeta(x)+\zeta(\nu).
\label{defzeta}
\end{equation}
Thus
\begin{eqnarray*}
{\biggl| \matrix{\Phi(x;\nu)&\Phi(y;\nu)\cr\Phi(x;\nu)\sp\prime&
\Phi(y;\nu)\sp\prime\cr}\biggr| }
&=&{\Phi(x;\nu)\Phi(y;\nu)\biggl[ \bigr(\ln \Phi(y;\nu)\bigl)\sp\prime
-\bigr(\ln \Phi(x;\nu)\bigl)\sp\prime\biggr] } \hfill\\
&=&\Phi(x;\nu)\Phi(y;\nu)\biggl[
\zeta(\nu-x)+\zeta(x)+\zeta(-y)+\zeta(y-\nu)\biggr]. \hfill
\end{eqnarray*}
Upon using the definition of $\Phi$, the right hand side of this
equation takes the form
\begin{equation}
\Phi(x+y;\nu){\sigma(\nu-x)\sigma(\nu-y)\sigma(x+y)\over
\sigma(\nu-x-y)\sigma(\nu)\sigma(x)\sigma(y) }
\biggl[ \zeta(\nu-x)+\zeta(x)+\zeta(-y)+\zeta(y-\nu)\biggr].
\label{rhs}
\end{equation}
After noting the two identities \cite{WW}
\begin{equation}
\zeta(x)+\zeta(y)+\zeta(z)-\zeta(x+y+z)=
{\sigma(x+y)\sigma(y+z)\sigma(z+x)\over\sigma(x)\sigma(y)\sigma(z)\sigma(x+y+z)}
\label{eq:zetas}
\end{equation}
and
\begin{equation}
\wp (x)-\wp(y)=
{\sigma(y-x)\sigma(y+x)\over\sigma\sp2(y)\sigma\sp2(x)}
\label{eq:wps}
\end{equation}
we find (\ref{rhs}) simplifies to
$\Phi(x+y;\nu)\bigl[{\wp }(x)-{\wp }(y)\bigr] $,
where ${\wp }(x)=-\zeta\sp\prime(x)$ is the Weierstrass ${\wp }$-function.
Putting these together yields the addition formula
\begin{equation}
\Phi(x+y;\nu)=
{{\biggl| \matrix{\Phi(x;\nu)&\Phi(y;\nu)\cr\Phi(x;\nu)\sp\prime&
\Phi(y;\nu)\sp\prime\cr}\biggr| }\over
{\wp }(x)-{\wp }(y) },
\label{addn}
\end{equation}
and consequently a solution of (\ref{functional})\ with the stated form.
Further, from (\ref{addn}) we see the solution to (\ref{calfun})
mentioned in the introduction.
The general solution (\ref{soln}) involves the two nonzero constants
$\nu_1$ and $\nu_2$.
Let us see how our group of symmetries enables $\phi_1(x)=\Phi(x;\nu_1)$
to occur as a limit $\nu_2\rightarrow0$.
Consider the ${\cal G}$ action on the general solution $\phi_i(x)$ of
theorem 1 given by $\phi_i(x)\rightarrow \tilde\phi_i(x)$ where
\begin{eqnarray*}
\Biggl( \tilde\phi_1(x), {\tilde\phi_2(x)\choose\tilde\phi_3(x)},&&
{\tilde\phi_4(x)\choose\tilde\phi_5(x)}
\Biggr) =\\ \\
&& \Biggl( {{ e\sp{\zeta(\nu_2) x}\over-\nu_2} \phi_1(x),
{ \Phi(x;\nu_1)\choose \Phi\sp\prime(x;\nu_1)},
{e\sp{-\zeta(\nu_2)x} \Phi(x;\nu_2)\choose
-\nu_2\,e\sp{-\zeta(\nu_2)x} \Phi\sp\prime(x;\nu_2)} \Biggr)}.
\end{eqnarray*}
Now
$$
\lim_{\nu_2\rightarrow0}\tilde\phi_1(x)=\Phi(x;\nu_1),
$$
and
$$
\lim_{\nu_2\rightarrow0}
{ \biggl|\matrix{\tilde\phi_4(x)&\tilde\phi_4(y)\cr
\tilde\phi_5(x)&\tilde\phi_5(y)\cr}\biggr|}
={\wp }(x)-{\wp }(y) .
$$
Thus (\ref{calfun}) arises as the $\nu_2\rightarrow0$ of (\ref{functional}).
\section{Proof of Theorem 1}
Our proof of Theorem 1 proceeds in two stages. First we will use the
symmetry (\ref{symm}) to transform (\ref{functional}) into a particularly
simple canonical form. This form may be immediately integrated to
yield a functional
equation studied by Bruschi and Calogero \cite{BCb}; by appealing to
their result our Theorem 1 will follow. The first stage of this process
is entirely algorithmic and consequently we may readily identify the
parameters that appear in our solution. We begin with
\begin{defn}A point $x_0\in\mathbb{C}$ is said to be {\it generic} for
(\ref{functional}) if
\begin{itemize}
\item[1)] $\phi_k(x)$ is regular at $x_0$ for $k=2\ldots5$,
\item[2)] $\phi_1(x)$ is regular at $2x_0$,
\item[3)]$ { { \biggl|
\matrix{\phi_2(x_0)&\phi_2\sp\prime(x_0)\cr
\phi_3(x_0)&\phi_3\sp\prime(x_0)\cr}\biggr|}
\ne0\quad\quad { \biggl|
\matrix{\phi_4(x_0)&\phi_4\sp\prime(x_0)\cr
\phi_5(x_0)&\phi_5\sp\prime(x_0)\cr}\biggr|} }\ne0.$
\end{itemize}
\end{defn}
Now let $x_0$ be a generic point. Using at first the matrices $U$ and $V$
of transformation (\ref{symm})
we may choose linear combinations of $\phi_k$ ($k:2\ldots5$) such that
(\ref{functional}) becomes
$$
\tilde\phi_1(x+y)=
{ { \biggl|\matrix{\tilde\phi_2(x)&\tilde\phi_2(y)\cr\tilde\phi_3(x)&
\tilde\phi_3(y)\cr}\biggr|}
\over
{ \biggl|\matrix{\tilde\phi_4(x)&\tilde\phi_4(y)\cr\tilde\phi_5(x)&
\tilde\phi_5(y)\cr}\biggr|} },
$$
and (for $k=1,2$)
\begin{equation}
\tilde\phi_{2k}(0)=\tilde\phi_{2k+1}\sp\prime(0)=0,\quad\quad
\tilde\phi_{2k}\sp\prime(0)=\tilde\phi_{2k+1}(0)=1.
\label{init}
\end{equation}
The arguments of the functions have been shifted to be centred on
$x_0$ (or $2x_0$ in the case of $\tilde\phi_1(x)$).
Here we have set
$$
\tilde\phi_1(x)=c\, \phi_1(x+2 x_0)
$$
and (for $k=1,2$)
$$
\Biggl(
\matrix{\tilde\phi_{2k}(x)\cr \tilde\phi_{2k+1}(x)\cr}\Biggr)=
\Biggl(
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\Biggr)\sp{-1}
\Biggl(
\matrix{\phi_{2k}(x+x_0)\cr \phi_{2k+1}(x+x_0)\cr}
\Biggr).
$$
The constant $c$ here is just the ratio of the appropriate determinants
specified in (\ref{symm}). We next observe
\begin{lem}
For $k=1,2$ we may write
$$
\Biggl(\matrix{\tilde\phi_{2k}(x)\cr \tilde\phi_{2k+1}(x)\cr}\Biggr)=
{1\over\gamma_k(x)}
\Biggl(\matrix{\xi_{k}(x)\cr \xi_{k}\sp\prime(x)\cr}\Biggr),
$$
where $\gamma_k(x), \xi_{k}(x)$ are regular at $0$ and
$$
\xi_{k}(0)=0,\quad\quad \xi_{k}\sp\prime(0)=\gamma_k(0)=1.
$$
Further, upon writing
$\xi_{k}(x)=e\sp{\lambda_k x}\tilde\xi_{k}(x)$
with $\lambda_k=-\tilde\phi_{2k}\sp{\prime\prime}(0)/2$
the function $\tilde\xi_{k}(x)$, regular at $0$, satisfies
$$
\tilde\xi_{k}(0)=\tilde\xi_{k}\sp{\prime\prime}(0)=0,\quad\quad
\tilde\xi_{k}\sp\prime(0)=1.
$$
\end{lem}
\begin{proof}
Upon differentiating $\xi_{k}(x)=\gamma_k(x)\tilde\phi_{2k}(x)$ and
comparing with
$\xi_{k}\sp\prime(x)=\gamma_k(x)\tilde\phi_{2k+1}(x)$ we see that
\begin{equation}
{\gamma_k\sp\prime(x)\over \gamma_k(x)}=
{ \tilde\phi_{2k+1}(x)-\tilde\phi_{2k}\sp\prime(x)\over \tilde\phi_{2k}(x)}.
\label{gammak}
\end{equation}
The only issue is whether the righthand side of this differential equation
is regular at $x=0$. Using (\ref{init}) and l'H\^opital's rule
we find
$$
{\gamma_k\sp\prime(0)\over \gamma_k(0)}=
{ \tilde\phi_{2k+1}\sp\prime(0)-\tilde\phi_{2k}\sp{\prime\prime}(0)
\over \tilde\phi_{2k}\sp\prime(0)}=-\tilde\phi_{2k}\sp{\prime\prime}(0).
$$
and so $\gamma_k(x)$ and hence $\xi_{k}(x)$ are regular at $0$.
Now $\gamma_k(x)$ is determined by (\ref{gammak}) given an initial
condition which we choose to be $\gamma_k(0)=1$. The remaining
initial conditions for $\xi_{k}(x)$ follow from (\ref{init}).
Indeed from
$$
\tilde\phi_{2k+1}\sp{\prime}(x)=
{\xi_{k}\sp{\prime\prime}(x)\gamma_k(x)-\xi_{k}\sp{\prime}(x)
\gamma_k\sp\prime(x)\over \gamma_k\sp2(x)}
$$
we also find
$$
\xi_{k}\sp{\prime\prime}(0)=-\tilde\phi_{2k}\sp{\prime\prime}(0).
$$
Now upon writing $\xi_{k}(x)=e\sp{\lambda_k x}\tilde\xi_{k}(x)$
with $\lambda_k=-\tilde\phi_{2k}\sp{\prime\prime}(0)/2$
we obtain the final statement of the lemma.
\end{proof}
Thus far we havent used the exponential part of the symmetry
(\ref{symm}). Utilising this symmetry we set
$\tilde\xi_{0}(x)=e\sp{(\lambda_1-\lambda_2)x}\tilde\phi_1(x)$,
$\gamma(x)= e\sp{2(\lambda_1-\lambda_2)x}\gamma_2(x)/\gamma_1(x)$.
This scaling has (upon noting $2\lambda_k=\gamma_k\sp\prime(0)$)
the effect of making $\gamma\sp\prime(0)=0$. Thus we obtain
\begin{col}
At any generic point we may rewrite (\ref{functional}) using
the symmetry (\ref{symm}) as
\begin{equation}
\tilde\xi_{0}(x+y)=\gamma(x)\gamma(y)
{ { \biggl|\matrix{\tilde\xi_1(x)&\tilde\xi_1(y)\cr\tilde\xi_1\sp\prime(x)&
\tilde\xi_1\sp\prime(y)\cr}\biggr|}
\over
{ \biggl|\matrix{\tilde\xi_2(x)&\tilde\xi_2(y)\cr\tilde\xi_2\sp\prime(x)&
\tilde\xi_2\sp\prime(y)\cr}\biggr|} },
\label{revfunctional}
\end{equation}
where for $k=1,2$
\begin{equation}
\tilde\xi_{k}(0)=\tilde\xi_{k}\sp{\prime\prime}(0)=\gamma\sp\prime(0)=0,
\quad\quad \tilde\xi_{k}\sp\prime(0)= \gamma(0)=1.
\label{initials}
\end{equation}
\end{col}
Given the complexity of the differential equation (\ref{gammak}) one may
wonder whether (\ref{revfunctional}) simplifies much further. In fact we
find
\begin{lem}
The functional equation (\ref{revfunctional}), and
consequently (\ref{functional}), may be written as
\begin{equation}
\partial\biggl({\tilde\xi_1(x+y)\over\tilde\xi_1(x)\tilde\xi_1(y)}\biggl)
=
\partial\biggl({\tilde\xi_2(x+y)\over\tilde\xi_2(x)\tilde\xi_2(y)}\biggl),
\label{newfunctional}
\end{equation}
where $\partial=\partial_x-\partial_y$. Further
\begin{equation}
\tilde\xi_{0}(x)={\tilde\xi_2(x)\over\tilde\xi_1(x)}.
\label{eqnsub}
\end{equation}
\end{lem}
\begin{proof}
Taking the logarithmic derivative of (\ref{revfunctional}) with respect to
$\partial=\partial_x-\partial_y$ we obtain
\begin{equation}
0={\gamma\sp\prime(x)\over\gamma(x)}-{\gamma\sp\prime(y)\over\gamma(y)}
+{\partial\sp2\left(\tilde\xi_1(x)\tilde\xi_1(y)\right)\over\partial
\left(\tilde\xi_1(x)\tilde\xi_1(y)\right)}
-{\partial\sp2\left(\tilde\xi_2(x)\tilde\xi_2(y)\right)\over\partial
\left(\tilde\xi_2(x)\tilde\xi_2(y)\right)}.
\label{intermed1}
\end{equation}
Now employing (\ref{initials}) one finds
$$
\partial\left(\tilde\xi_k(x)\tilde\xi_k(y)\right)\vert_{y=0}
=\tilde\xi_k\sp\prime(x)\tilde\xi_k(y)-\tilde\xi_k(x)\tilde\xi_k\sp\prime(y)
\vert_{y=0}=-\tilde\xi_k(x)
$$
and similarly
$$
\partial\sp2\left(\tilde\xi_k(x)\tilde\xi_k(y)\right)\vert_{y=0}=
-2\tilde\xi_k\sp\prime(x).
$$
Upon setting $y=0$ in (\ref{intermed1}) and with these simplifications
we obtain the differential equation
$$
0={\gamma\sp\prime(x)\over\gamma(x)}+{
2\tilde\xi_1\sp\prime(x)\over\tilde\xi_1(x)}-
{2\tilde\xi_2\sp\prime(x)\over\tilde\xi_2(x)}
$$
with solution
\begin{equation}
\gamma(x)=c\, {\tilde\xi_2\sp2(x)\over\tilde\xi_1\sp2(x)}.
\label{gammadef}
\end{equation}
Again using l'H\^opital's rule and (\ref{initials}) we find the
constant $c=1$.
Therefore (\ref{revfunctional}) may be rewritten as
\begin{equation}
\tilde\xi_{0}(x+y)=
{\tilde\xi_2\sp2(x)\over\tilde\xi_1\sp2(x)}
{\tilde\xi_2\sp2(y)\over\tilde\xi_1\sp2(y)}
{ { \Biggl|\matrix{\tilde\xi_1(x)&\tilde\xi_1(y)\cr\tilde\xi_1\sp\prime(x)&
\tilde\xi_1\sp\prime(y)\cr}\Biggr|}
\over
{ \Biggl|\matrix{\tilde\xi_2(x)&\tilde\xi_2(y)\cr\tilde\xi_2\sp\prime(x)&
\tilde\xi_2\sp\prime(y)\cr}\Biggr|} }
=
{ { \Biggl|\matrix{{1\over\tilde\xi_1(x)}&{1\over\tilde\xi_1(y)}\cr
\big({1\over\tilde\xi_1(x)}\big)\sp\prime&
\big({1\over\tilde\xi_1(y)}\big)\sp\prime\cr}\Biggr|}
\over
{ \Biggl|\matrix{{1\over\tilde\xi_2(x)}&{1\over\tilde\xi_2(y)}\cr
\big({1\over\tilde\xi_2(x)}\big)\sp\prime&
\big({1\over\tilde\xi_2(y)}\big)\sp\prime\cr}\Biggr|}
}
=
{\partial\big({1\over\tilde\xi_1(x)\tilde\xi_1(y)}\big) \over
\partial\big({1\over\tilde\xi_2(x)\tilde\xi_2(y)}\big)}.
\label{intermed2}
\end{equation}
Letting $y\rightarrow0$ we find
$$
\tilde\xi_{0}(x)={\tilde\xi_2(x)\over\tilde\xi_1(x)}
$$
as required. Utilizing (\ref{eqnsub}) we may immediately rewrite
(\ref{intermed2}) in the stated form (\ref{newfunctional}).
\end{proof}
We observe that at this stage the symmetry (\ref{symm}) has enabled us
to transform (\ref{functional}) into the form specified by theorem 1.
The solution will follow once we show $1/\tilde\xi_k(x)=\Phi(x;\nu_k)$.
Now (\ref{newfunctional}) may be immediately integrated to give
$$
{\tilde\xi_1(x+y)\over\tilde\xi_1(x)\tilde\xi_1(y)}
={\tilde\xi_2(x+y)\over\tilde\xi_2(x)\tilde\xi_2(y)} +\Theta(x+y).
$$
Upon setting $\alpha(x)={\tilde\xi_2(x)/\tilde\xi_1(x)}$ and
$\psi(x)=\Theta(x)/ \tilde\xi_2(x)$ this may be rearranged into the form
\begin{equation}
{\alpha(x+y)\over\alpha(x)\alpha(y)}=
1+\tilde\xi_2(x)\tilde\xi_2(y)\psi(x+y)
\label{bcalfun}
\end{equation}
which is the functional equation studied by Bruschi and Calogero
\cite{BCb}.
Calling upon the general analytic solution obtained by these authors,
together with our initial conditions (\ref{initials}), we find\footnote{
For example, from \cite{BCb} we obtain
$\tilde\xi_2(x)=A e\sp{cx}\sigma(ax\vert\omega,\omega\sp\prime)/
\sigma(ax+\nu\vert\omega,\omega\sp\prime)$. Using the property
$\sigma(ax\vert a\omega,a\omega\sp\prime)=
a\sigma(ax\vert\omega,\omega\sp\prime)$ and the definition of $\Phi(x;\nu)$
we may rewrite this as
$\tilde\xi_2(x)=(A/\sigma(\nu/a)) e\sp{(c-\zeta(\nu/a))x} /\Phi(z;-\nu/a)$.
Now the $x\rightarrow 0$ limit shows
$(A/\sigma(\nu/a)) e\sp{(c-\zeta(\nu/a))x}=1$.}
$1/\tilde\xi_k(x)=\Phi(x;\nu_k)$ as required.
\section{Proof of Theorem 2}
It is useful at the outset to gather together the various transformations
introduced in the last section:
\begin{eqnarray}
\Biggl(
\matrix{\tilde\phi_{2k}(x)\cr \tilde\phi_{2k+1}(x)\cr}\Biggr)&=&
\Biggl(
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\Biggr)\sp{-1}
\Biggl(
\matrix{\phi_{2k}(x+x_0)\cr \phi_{2k+1}(x+x_0)\cr}
\Biggr) \label{transform1}\\
&=&{1\over\gamma_k(x)}
\Biggl(\matrix{\xi_{k}(x)\cr \xi_{k}\sp\prime(x)\cr}\Biggr)
={e\sp{\lambda_k x}\over\gamma_k(x)}
\Biggl( \matrix{1&0\cr \lambda_k&1\cr}\Biggr)
\Biggl( \matrix{\tilde\xi_k(x)\cr \tilde\xi_k\sp\prime(x)\cr}\Biggr)
\label{transform2}\\
&=&{e\sp{\lambda_k x}\over\gamma_k(x)\Phi\sp2(x;\nu_k)}
\Biggl( \matrix{1&0\cr \lambda_k&-1\cr}\Biggr)
\Biggl( \matrix{\Phi(x;\nu_k)\cr \Phi\sp\prime(x;\nu_k)\cr}\Biggr)
\label{transform3}\\
\tilde\xi_0(x)&=&e\sp{(\lambda_1-\lambda_2) x}\,
{\phi_1(x+2 x_0)\over\phi_1(2 x_0)}={\Phi(x;\nu_1)\over\Phi(x;\nu_2)}.
\label{transform4}
\end{eqnarray}
Let us introduce the function
\begin{equation}
f(x)= \gamma_k(x)\Phi\sp2(x;\nu_k)e\sp{-2\lambda_k x}.
\label{transform5}
\end{equation}
Observe that (\ref{gammadef}) entails the function
$f(x)$ is independent of $k$,
$$
\gamma_1(x)\Phi\sp2(x;\nu_1)e\sp{-2\lambda_1 x}
=
\gamma_2(x)\Phi\sp2(x;\nu_k)e\sp{-2\lambda_k x}.
$$
With this definition we may rewrite (\ref{transform1}) and
(\ref{transform3}) to give
\begin{eqnarray}
\left( \matrix{\phi_{2k}(x+x_0)\cr \phi_{2k+1}(x+x_0)\cr} \right)
&=&
{e\sp{-\lambda_k x}\over f(x)}
\Biggl(
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\Biggr)
\Biggl( \matrix{1&0\cr \lambda_k&-1\cr}\Biggr)
\Biggl( \matrix{\Phi(x;\nu_k)\cr \Phi\sp\prime(x;\nu_k)\cr}\Biggr)
\nonumber
\\
\label{transform6}
\\
&=&
{e\sp{-\lambda_k x}\over f(x)}
\left(
\matrix{ \left|
\matrix{\Phi(x;\nu_k)&\phi_{2k}(x_0)\cr
\Phi\sp\prime(x;\nu_k)&\phi_{2k}\sp\prime(x_0)+\lambda_k\phi_{2k}(x_0)\cr}
\right|
\cr
\left|
\matrix{\Phi(x;\nu_k)&\phi_{2k+1}(x_0)\cr
\Phi\sp\prime(x;\nu_k)&\phi_{2k+1}\sp\prime(x_0)+\lambda_k\phi_{2k+1}(x_0)\cr}
\right|
\cr
}
\right)
\nonumber
\end{eqnarray}
and
\begin{equation}
\Biggl( \matrix{\Phi(x;\nu_k)\cr \Phi\sp\prime(x;\nu_k)\cr}\Biggr)
=
f(x)e\sp{\lambda_k x}\Biggl( \matrix{1&0\cr \lambda_k&-1\cr}\Biggr)
\Biggl(
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\Biggr)\sp{-1}
\Biggl(
\matrix{\phi_{2k}(x+x_0)\cr \phi_{2k+1}(x+x_0)\cr}
\Biggr)
\label{transform7}
\end{equation}
$$
\quad=
{f(x)e\sp{\lambda_k x}\over
\left|
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\right|
}\,
\left(
\matrix{ \left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(x_0)\cr}
\right|
\cr
\left|
\matrix{\phi_{2k}(x+x_0)&
\phi_{2k}\sp\prime(x_0)+\lambda_k\phi_{2k}(x_0)\cr
\phi_{2k+1}(x+x_0)&
\phi_{2k+1}\sp\prime(x_0)+\lambda_k\phi_{2k+1}(x_0)\cr}
\right|
\cr
}
\right).
\nonumber
$$
Now (\ref{transform4}) and (\ref{transform6}) are of the form stated
in theorem 2 provided we can show $f(x)$, defined in (\ref{transform5}),
can also be put into the form of the theorem. To see this
note that (\ref{transform1}) shows
\begin{eqnarray}
\tilde\phi_{2k}(x)&=&
{ {\phi_{2k+1}(x_0)\phi_{2k}(x+x_0)- \phi_{2k}(x_0)\phi_{2k+1}(x+x_0)}\over
{\phi_{2k+1}(x_0)\phi_{2k}\sp\prime(x_0)-
\phi_{2k}(x_0)\phi_{2k+1}\sp\prime(x_0)}
}
\nonumber
\\
\label{transform8}
\\
&=&
{
\left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(x_0)\cr}
\right|
\over
\left|
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\right|
}
\nonumber
\end{eqnarray}
while from (\ref{transform3}) we see
\begin{equation}
\gamma_k(x)={ e\sp{\lambda_k x}\over \Phi(x;\nu_k)\tilde\phi_{2k}(x)}.
\label{transform9}
\end{equation}
Combining these thus shows
\begin{equation}
f(x)=
{e\sp{-\lambda_k x}\over \Phi(x;\nu_k)}\,
{
\left|
\matrix{\phi_{2k}\sp\prime(x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x_0)&\phi_{2k+1}(x_0)\cr}
\right|
\over
\left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(x_0)\cr}
\right|
}
\label{transform10}
\end{equation}
as required. Also from (\ref{transform8}) and the definition
$\lambda_k=-\tilde\phi_{2k}\sp{\prime\prime}(0)/2$ we find
\begin{eqnarray}
-2\lambda_k &=&
{ {\phi_{2k+1}(x_0)\phi_{2k}\sp{\prime\prime}(x_0)-
\phi_{2k}(x_0)\phi_{2k+1}\sp{\prime\prime}(x_0)}\over
{\phi_{2k+1}(x_0)\phi_{2k}\sp\prime(x_0)-
\phi_{2k}(x_0)\phi_{2k+1}\sp\prime(x_0)}
}
\label{transform11}\\
&=&
\partial_x\ln
{ \Biggl|\matrix{ \phi_{2k}\sp\prime(x+x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}\sp\prime(x+x_0)&\phi_{2k+1}(x_0)\cr}
\Biggl| }_{x=0}.
\label{transform12}
\end{eqnarray}
At this stage then we see that if we can determine $\Phi(x;\nu_k)$
all of the terms in (\ref{transform1}-\ref{transform4}) are determined
and we obtain the stated expressions for $\phi_1(x)$, $\phi_2(x)$,
$\phi_3(x)$, $\phi_4(x)$, $\phi_5(x)$ and $f(x)$ given in theorem 2.
It remains therefore to determine the parameters $g_2$, $g_3$
specifying the elliptic functions $\Phi(x;\nu_k)$ as well as
$\nu_1$, $\nu_2$.
To this end we utilise (\ref{transform7}) to give
\begin{eqnarray*}
{\Phi\sp\prime(x;\nu_k) \over \Phi(x;\nu_k)}-\lambda_k
&=&
{
\left|
\matrix{\phi_{2k}(x+x_0)& \phi_{2k}\sp\prime(x_0)\cr
\phi_{2k+1}(x+x_0)& \phi_{2k+1}\sp\prime(x_0)\cr}
\right|
\over
\left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(x_0)\cr}
\right|
}
\\
&=&
\partial_y \ln
\left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(y+x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(y+x_0)\cr}
\right|
\Biggl|_{y=0}.
\end{eqnarray*}
Upon using (\ref{defzeta}) to simplify the left-hand side of this
equality we thus obtain the first equality of theorem 2,
\begin{equation}
\partial_y \ln
\left|
\matrix{\phi_{2k}(x+x_0)&\phi_{2k}(y+x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(y+x_0)\cr}
\right|
\Biggl|_{y=0}
=
\zeta(\nu_k)-\zeta(x)-\zeta(\nu_k-x)-\lambda_k,
\label{transform13}
\end{equation}
and consequently
\begin{equation}
\partial_x\partial_y\ln
{ \Biggl|\matrix{ \phi_{2k}(x+x_0)&\phi_{2k}(y+x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(y+x_0)\cr}
\Biggr|_{y=0} }
\label{eq:Fdef}
=\wp(x)-\wp(\nu_k-x).
\end{equation}
In fact we have the more general result
\begin{eqnarray*}
&\partial_x\partial_y&\ln
{ \Biggl|\matrix{ \phi_{2k}(x+x_0)&\phi_{2k}(y+x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(y+x_0)\cr}
\Biggr| }
=
\partial_x\partial_y\ln
{ \Biggl|\matrix{ \Phi(x;\nu_k)&\Phi(y;\nu_k)\cr
\Phi\sp\prime(x;\nu_k)&\Phi\sp\prime(y;\nu_k)\cr}
\Biggr| }
\\
&=&
\partial_x\partial_y\ln
\biggl[\Phi(x+y;\nu_k)\Bigl( {\wp }(x)-{\wp }(y) \Bigr)\biggr]
\\
&=&
\wp(x+y)-\wp(\nu_k-x-y)+
{ \wp\sp\prime(x)\wp\sp\prime(y)\over \Bigl( {\wp }(x)-{\wp }(y) \Bigr)\sp2}
\end{eqnarray*}
from which (\ref{eq:Fdef}) arises as the $y\rightarrow0$ limit.
It remains to how that the Laurent series
of (\ref{transform13}) (or equivalently of (\ref{eq:Fdef})) determines the
parameters of $\Phi(x;\nu_k)$. Set
\begin{equation}
\zeta(\nu_k)-\zeta(x)-\zeta(\nu_k-x)-\lambda_k
=
-{1\over x}-\lambda_k +\sum_{l=0}F_l\, {x\sp{l+1}\over (l+1)!}
\label{transform15}
\end{equation}
or equivalently
\begin{equation}
\wp(x)-\wp(\nu_k-x)=
{1\over x\sp2}+\sum_{l=0}{F_l\over l!}\, x\sp{l}.
\label{transform16}
\end{equation}
While the coefficients $F_l$ in these expansions depend on
$k=1,2$ we will avoid including this in our notation: certainly the
combinations of these coefficients giving $g_2$ and $g_3$
are independent of $k$.
Now the left-hand side of (\ref{transform15}) has the expansion
$$
-{1\over x}-\lambda_k -\wp(\nu_k)\, x+\wp\sp{\prime}(\nu_k)\,{x\sp2\over2}+
(2c_2 -\wp\sp{\prime\prime}(\nu_k)){x\sp3\over 3!}+\ldots
$$
while that of (\ref{transform16}) begins
$$
{1\over x\sp2}+c_2\,x\sp2+c_3\,x\sp4+\ldots-\Big\lbrace
\wp(\nu_k)-x\wp\sp{\prime}(\nu_k)+{x\sp2 \over2}\wp\sp{\prime\prime}(\nu_k)+
\ldots\Big\rbrace
$$
From either of these we see
$$
F_0=-\wp(\nu_k),\quad F_1=\wp\sp{\prime}(\nu_k),\quad
F_2=2c_2 -\wp\sp{\prime\prime}(\nu_k),
$$
whereupon utilising (\ref{eq:wpdef}) we obtain
\begin{equation}
c_2={F_2+6 F_0\sp2 \over 12}={g_2\over20}
\quad{\rm and}\quad
g_3= 6 F_0\sp3 -F_1\sp2 +{5\over3}F_0 F_2.
\label{transform18}
\end{equation}
Thus, as stated in Theorem 2, we may obtain the parameters of the elliptic
functions from the Laurent expansion (\ref{transform15}) for either
choice of $k$, the combinations of the coefficients in (\ref{transform18})
being independent of $k$. The constant terms in the two expansions
the determine $\nu_1$, $\nu_2$ via $F_0=-\wp(\nu_k;g_2,g_3)$.
We now have now established all of Theorem 2. It is perhaps useful to
conclude the section with a lemma that implements the theorem.
\begin{lem}
Let
\begin{eqnarray*}
\partial_x\partial_y\ln
{ \Biggl|\matrix{ \phi_{2k}(x+x_0)&\phi_{2k}(y+x_0)\cr
\phi_{2k+1}(x+x_0)&\phi_{2k+1}(y+x_0)\cr}
\Biggr|_{y=0} }
&=& -{1\over x}-\lambda_k +\sum_{l=0}F_l\, {x\sp{l+1}\over (l+1)!}
\\
&=&
{ h\sp{\prime}(0)h\sp{\prime}(x)\over (h(x)-h(0))\sp2 }
\end{eqnarray*}
where $\displaystyle{ h(x)={\phi_{2k}(x+x_0)\over \phi_{2k+1}(x+x_0)} }$.
Set $\displaystyle{h_k= {h\sp{(k+1)}(0)\over{ (k+1)!\, h\sp{\prime}(0)}} }$.
Then
$$
F_0= - \left|\matrix{h_1&1\cr h_2&h_1\cr}\right|,\quad
F_1= 2 \left|\matrix{h_1&1&0\cr h_2&h_1&1\cr h_3&h_2&h_1\cr}\right|,\quad
F_2=-6 \left|\matrix{h_1&1&0&0\cr h_2&h_1&1&0\cr h_3&h_2&h_1&1\cr
h_4&h_3&h_2&h_1\cr}\right|.
$$
\end{lem}
\section{Examples}
We shall now consider the classical addition theorems of
the Jacobi elliptic functions and then the functional equations
(\ref{dexponential}) and (\ref{eq:biggy}) as examples of our theory.
We have collected several standard results pertaining to
elliptic functions that are of use in our computations in Appendix A.
\noindent{\bf Example 1.}
As a first application of our theory we consider the addition
theorems for the Jacobi elliptic functions $\mathop{\rm dn}\nolimits(x)$, $\mathop{\rm cn}\nolimits(x)$ and
$\mathop{\rm sn}\nolimits(x)$ where $\mathop{\rm dn}\nolimits(x)\equiv \mathop{\rm dn}\nolimits(x|m)$ and so on. These may be
cast in the form of (\ref{functional}) as
$$
\mathop{\rm dn}\nolimits(x+y)={
\left|\matrix{\mathop{\rm cn}\nolimits\sp\prime(x)&\mathop{\rm cn}\nolimits\sp\prime(y) \cr
\mathop{\rm cn}\nolimits(x)&\mathop{\rm cn}\nolimits(y)\cr}\right|
\over
\left|\matrix{\mathop{\rm sn}\nolimits\sp\prime(x)&\mathop{\rm sn}\nolimits\sp\prime(y) \cr
\mathop{\rm sn}\nolimits(x)&\mathop{\rm sn}\nolimits(y)\cr}\right|
}\quad{\rm (Jacobi)},\quad\quad
\mathop{\rm cn}\nolimits(x+y)={1\over k\sp2}{
\left|\matrix{\mathop{\rm dn}\nolimits\sp\prime(x)&\mathop{\rm dn}\nolimits\sp\prime(y) \cr
\mathop{\rm dn}\nolimits(x)&\mathop{\rm dn}\nolimits(y)\cr}\right|
\over
\left|\matrix{\mathop{\rm sn}\nolimits\sp\prime(x)&\mathop{\rm sn}\nolimits\sp\prime(y) \cr
\mathop{\rm sn}\nolimits(x)&\mathop{\rm sn}\nolimits(y)\cr}\right|
}
$$
and
$$
\mathop{\rm sn}\nolimits(x+y)={
\left|\matrix{1&1\cr \mathop{\rm sn}\nolimits\sp2(x)&\mathop{\rm sn}\nolimits\sp2(y)\cr}\right|
\over
\left|\matrix{\mathop{\rm sn}\nolimits\sp\prime(x)&\mathop{\rm sn}\nolimits\sp\prime(y) \cr
\mathop{\rm sn}\nolimits(x)&\mathop{\rm sn}\nolimits(y)\cr}\right|
}\quad{\rm (Cayley)}.
$$
Let us now apply our theorem to the first equality.
The first step is to choose an appropriate generic point $x_0$.
This means we wish $x_0$ to be a regular point for $\mathop{\rm cn}\nolimits(x)$, $\mathop{\rm sn}\nolimits(x)$
and $\mathop{\rm dn}\nolimits(2x)$ as well as
\begin{eqnarray*}
0&\ne& \left|\matrix{\mathop{\rm cn}\nolimits\sp\prime(x_0)&\mathop{\rm cn}\nolimits\sp{\prime\prime}(x_0)\cr
\mathop{\rm cn}\nolimits(x_0)&\mathop{\rm cn}\nolimits\sp\prime(x_0)\cr}\right|=1-m+m\,\mathop{\rm cn}\nolimits\sp4(x_0),
\\
0&\ne& \left|\matrix{\mathop{\rm sn}\nolimits\sp\prime(x_0)&\mathop{\rm sn}\nolimits\sp{\prime\prime}(x_0)\cr
\mathop{\rm sn}\nolimits(x_0)&\mathop{\rm sn}\nolimits\sp\prime(x_0)\cr}\right|=1-m\,\mathop{\rm sn}\nolimits\sp4(x_0).
\end{eqnarray*}
Thus we can take $x_0=0$ for this example.
Using (\ref{transform11}) we find
$$
-2\lambda_1=\partial_x\ln\mathop{\rm cn}\nolimits\sp\prime(x)\big\vert_{x=0} =0\quad{\rm and}\quad
-2\lambda_2=\partial_x\ln\mathop{\rm sn}\nolimits\sp\prime(x)\big\vert_{x=0} =0.
$$
Further, with $h(x)=\phi_2(x)/\phi_3(x)=\mathop{\rm cn}\nolimits\sp\prime(x)/\mathop{\rm cn}\nolimits(x)$ we obtain
$$
F(x)={1-m+m\,\mathop{\rm cn}\nolimits\sp4(x)\over \mathop{\rm sn}\nolimits\sp2(x)\mathop{\rm dn}\nolimits\sp2(x)}
={1\over x\sp2}+ {1-2 m\over3}+{1+14 m-14 m\sp2 \over 15}x\sp2+\ldots
$$
while with $h(x)=\phi_4(x)/\phi_5(x)=\mathop{\rm sn}\nolimits\sp\prime(x)/\mathop{\rm sn}\nolimits(x)$ we obtain
$$
F(x)={1-m\,\mathop{\rm sn}\nolimits\sp4(x)\over \mathop{\rm sn}\nolimits\sp2(x)}
={1\over x\sp2}+ {1+m\over3}+{1-16 m+m\sp2 \over 15}x\sp2+\ldots
$$
In both cases we find
$$
g_2={4\over3}(1-m+m\sp2)\quad{\rm and}\quad
g_3={4\over27}(m-2)(2m-1)(m+1),
$$
(the required equality providing a nontrivial check) which means
$$
e_1={2-m\over3},\quad e_2={2 m-1\over3}\quad{\rm and}\quad e_3={-1-m\over3}.
$$
Further,
$$\wp(\nu_1)={2 m-1\over3}\quad{\rm and}\quad
\wp(\nu_2)={-1-m\over3}.
$$
Comparison with (\ref{ehoms}) and (\ref{jacgs})
shows $\omega=K(m)$, $\omega\sp\prime=iK\sp\prime(m)$,
$\nu_1=K(m)+iK\sp\prime(m)$ and $\nu_2=K\sp\prime(m)$.
We may also calculate $f(x)=1/\mathop{\rm sn}\nolimits\sp2(x)$ and upon
using (\ref{eq:phiJacs}) our identity may be rewritten as
\begin{eqnarray*}
\mathop{\rm dn}\nolimits(x+y)&=&{\Phi(x+y;K(m)+iK\sp\prime(m))\over \Phi(x+y;iK\sp\prime(m))} \\
&=&
{{\biggl| \matrix{\Phi(x;K(m)+iK\sp\prime(m))&\Phi(y;K(m)+iK\sp\prime(m))\cr
\Phi(x;K(m)+iK\sp\prime(m))\sp\prime&
\Phi(y;K(m)+iK\sp\prime(m))\sp\prime\cr}\biggr| }\over
{\biggl| \matrix{\Phi(x;iK\sp\prime(m))&\Phi(y;iK\sp\prime(m))\cr
\Phi(x;iK\sp\prime(m))\sp\prime&
\Phi(y;iK\sp\prime(m))\sp\prime\cr}\biggr| }
}\\
&=&
{{1\over\mathop{\rm sn}\nolimits\sp2(x)}
\left|\matrix{\mathop{\rm cn}\nolimits\sp\prime(x)&\mathop{\rm cn}\nolimits\sp\prime(y) \cr
\mathop{\rm cn}\nolimits(x)&\mathop{\rm cn}\nolimits(y)\cr}\right|
\over
{1\over\mathop{\rm sn}\nolimits\sp2(x)}
\left|\matrix{\mathop{\rm sn}\nolimits\sp\prime(x)&\mathop{\rm sn}\nolimits\sp\prime(y) \cr
\mathop{\rm sn}\nolimits(x)&\mathop{\rm sn}\nolimits(y)\cr}\right|
}\\
\end{eqnarray*}
The second identity may be treated in the same manner yielding
$\nu_1=K(m)$ and $\nu_2=K\sp\prime(m)$.
The third identity is a little different. It may be rewritten as
\begin{eqnarray*}
\Phi(x+y;iK\sp\prime(m))&=&
{1\over \mathop{\rm sn}\nolimits(x+y)}=
{ {1\over\mathop{\rm sn}\nolimits\sp2(x)}
\left|\matrix{\mathop{\rm sn}\nolimits\sp\prime(x)&\mathop{\rm sn}\nolimits\sp\prime(y) \cr
\mathop{\rm sn}\nolimits(x)&\mathop{\rm sn}\nolimits(y)\cr}\right|
\over
{1\over\mathop{\rm sn}\nolimits\sp2(x)}
\left|\matrix{1&1\cr \mathop{\rm sn}\nolimits\sp2(x)&\mathop{\rm sn}\nolimits\sp2(y)\cr}\right|
}\\
&=&
{ {\biggl| \matrix{\Phi(x;iK\sp\prime(m))&\Phi(y;iK\sp\prime(m))\cr
\Phi(x;iK\sp\prime(m))\sp\prime&
\Phi(y;iK\sp\prime(m))\sp\prime\cr}\biggr| }
\over
{\biggl| \matrix{\Phi\sp2(x;iK\sp\prime(m))&\Phi\sp2(y;iK\sp\prime(m))\cr
1&1\cr}\biggr| }
}.
\end{eqnarray*}
Now
$$
\Phi\sp2(x;iK\sp\prime(m))-\Phi\sp2(y;iK\sp\prime(m))
=\wp(x)-\wp(y)
$$
and the required identity follows from the general solution by the limiting
procedure described in section 3.
\noindent{\bf Example 2.}
We shall now determine the general analytic solution of
$$
\phi_1(x+y)= \phi_4(x)\phi_5(y)+\phi_4(y)\phi_5(x)=
{ { \biggl|\matrix{\phi_2(x)&\phi_2(y)\cr\phi_3(x)&\phi_3(y)\cr}\biggr|}
\over
{ \biggl|\matrix{\phi_4(x)&\phi_4(y)\cr\phi_5(x)&\phi_5(y)\cr}\biggr|} },
$$
where $ \phi_2(x)=\phi_4\sp2(x)$, $\phi_3(x)=\phi_5\sp2(x)$.
The particular case $\phi_1(x)=\phi_4(x)$ was treated in
\cite{BK}.
Suppose $x_0$ is a generic point. Then from
$$
0\ne
\biggl|
\matrix{\phi_2(x_0)&\phi_2\sp\prime(x_0)\cr
\phi_3(x_0)&\phi_3\sp\prime(x_0)\cr}
\biggr|
=
2 \phi_4(x_0) \phi_5(x_0)
\biggl|
\matrix{\phi_4(x_0)&\phi_4\sp\prime(x_0)\cr
\phi_5(x_0)&\phi_5\sp\prime(x_0)\cr}
\biggr|
$$
we see $\phi_4(x_0)\ne0$, $\phi_5(x_0)\ne0$ and
$\phi_1(2 x_0)= 2 \phi_4(x_0) \phi_5(x_0)\ne0$.
Further, from (\ref{transform11}), we find
\begin{equation}
\lambda_1=\lambda_2-{1\over2}
\left( {\phi_4\sp\prime(x_0)\over \phi_4(x_0)}
+ {\phi_5\sp\prime(x_0)\over \phi_5(x_0)}
\right).
\label{eq:lams}
\end{equation}
Our strategy is as follows. We will first determine
$\nu_1$, $\nu_2$, $\lambda_1$, $\lambda_2$, the parameters
describing the elliptic functions and the ratio
$\phi_4(x+x_0)\phi_5(x_0) / \phi_5(x+x_0)\phi_4(x_0)$.
Then from
\begin{eqnarray*}
\phi_1(x+2 x_0)
&=& \phi_4(x+x_0)\phi_5(x_0)+\phi_4(x_0)\phi_5(x+x_0) \\
&=&\phi_4(x+x_0)\phi_5(x_0)\left(
1+{\phi_5(x+x_0)\phi_4(x_0)\over \phi_4(x+x_0)\phi_5(x_0)}
\right) \\
&=&
e\sp{(\lambda_2-\lambda_1) x}\,{\Phi(x;\nu_1)\over\Phi(x;\nu_2)}
\phi_1(2 x_0)
\end{eqnarray*}
we will obtain
\begin{equation}
\phi_4(x+x_0)=
{2 \phi_4(x_0) e\sp{(\lambda_2-\lambda_1) x} \over
1+{\phi_5(x+x_0)\phi_4(x_0) / \phi_4(x+x_0)\phi_5(x_0) }
}
{\Phi(x;\nu_1)\over\Phi(x;\nu_2)}
\label{eq:strategy}
\end{equation}
with $\phi_5(x+x_0)$ immediately following.
Now from (\ref{transform6}) we obtain
\begin{equation}
{\phi_{2k}(x+x_0)\over \phi_{2k+1}(x+x_0)}
{\phi_{2k+1}(x_0)\over \phi_{2k}(x_0)}
=
1+ {N_k\over D_k}
\label{eq:NDks}
\end{equation}
where
$$
N_k= { \phi_{2k}\sp\prime(x_0)\over\phi_{2k}(x_0)}
-{ \phi_{2k+1}\sp\prime(x_0)\over\phi_{2k+1}(x_0)}
\quad\quad
D_k= { \phi_{2k+1}\sp\prime(x_0)\over\phi_{2k+1}(x_0)}+\lambda_k
-{\Phi\sp\prime(x;\nu_k)\over\Phi(x;\nu_k)}
$$
Here $N_1=2 N_2$ and by our assumption that $x_0$ was a generic point
these are non vanishing. Further, from
$ \phi_2(x)=\phi_4\sp2(x)$ and $\phi_3(x)=\phi_5\sp2(x)$, we see that
$$
1+N_1/D_1 =\left( 1+ N_2/D_2\right)\sp2.
$$
Expanding this shows $D_2\sp2=(D_2+N_2/2)D_1$ which upon using
(\ref{eq:lams}) yields
\begin{eqnarray}
\left(
{ \phi_5\sp\prime(x_0)\over\phi_5(x_0)} +\lambda_2
-{\Phi\sp\prime(x;\nu_2)\over\Phi(x;\nu_2)}
\right)\sp2
&=&
\left(
{ \phi_5\sp\prime(x_0)\over\phi_5(x_0)} +\lambda_2
+{N_2\over2}
-{\Phi\sp\prime(x;\nu_2)\over\Phi(x;\nu_2)}
\right)
\\ \nonumber &&\quad
\left(
{ \phi_5\sp\prime(x_0)\over\phi_5(x_0)} +\lambda_2
-{N_2\over2}
-{\Phi\sp\prime(x;\nu_1)\over\Phi(x;\nu_1)}
\right)
\label{eq:poles}
\end{eqnarray}
Suppose that $\nu_2$ is finite. Comparing the pole behaviour of each
side of (\ref{eq:poles}) shows that $\nu_1=\nu_2$ and consequently
that $N_2=0$, a contradiction.
The remaining possibility is that $\nu_2$ is infinite which
we now show to be a consistent solution. This can only
happen if the elliptic function degenerates into a hyperbolic or
trigonometric function and without loss of generality we choose the former.
In this case
\begin{equation}
\Phi(x;\nu)={\kappa\sinh \kappa(\nu-x)\over\sinh \kappa\nu
\sinh \kappa x}
e\sp{ x \kappa\coth\kappa\nu}\quad{\rm and}\quad
\Phi(x;\infty)={\kappa\over\sinh \kappa x}.
\label{eq:phiinff}
\end{equation}
Let us the suppose $\nu_2=\infty$.
Utilising (\ref{Phiinf}) and (\ref{Phiinfs}) we then must solve
$$
\left(
{ \phi_5\sp\prime(x_0)\over\phi_5(x_0)} +\lambda_2
+\kappa\, \coth\kappa x
\right)\sp2
=
\left(
{ \phi_5\sp\prime(x_0)\over\phi_5(x_0)} +\lambda_2
+{N_2\over2}
+\kappa\, \coth\kappa x
\right)\quad\quad\quad
$$
$$
\left(
{ \phi_5\sp\prime(x_0)\over\phi_5(x_0)} +\lambda_2
-{N_2\over2}
+\kappa\coth\kappa (\nu_1-x) +\kappa\coth\kappa x-\kappa\coth\kappa\nu_1
\right).
$$
This holds provided
$$
N_2\sp2= {4 \kappa\sp2\over\sinh \kappa\nu_1},
\quad\quad
{ \phi_5\sp\prime(x_0)\over\phi_5(x_0)} +\lambda_2
+{N_2\over2}
+\kappa\, \coth\kappa \nu_1=0
$$
which determines $\nu_1$, $\lambda_2$ and (via (\ref{eq:lams}))
$\lambda_1$ in terms of $\phi_{4}(x_0)$, $\phi_{5}(x_0)$,
$\phi_4\sp\prime(x_0)$ and $\phi_5\sp\prime(x_0)$.
The choice of sign in taking the square-root here is
arbitrary (just defining $\nu_1$) and we will take
$N_2 = -2\kappa/\sinh \kappa\nu_1$.
Substituting these into (\ref{eq:NDks}) we find
\begin{eqnarray*}
{\phi_{4}(x+x_0)\over \phi_{5}(x+x_0)}
{\phi_{5}(x_0)\over \phi_{4}(x_0)}
&=&
{\kappa\coth\kappa x- \kappa\, \coth\kappa \nu_1+{N_2/2}
\over \kappa\coth\kappa x- \kappa\, \coth\kappa \nu_1-{N_2/2}
}
\\ &=&
\coth(\kappa \nu_1/2) \tanh\kappa( \nu_1/2-x).
\end{eqnarray*}
Now employing (\ref{eq:phiinff}) shows
\begin{equation}
\phi_1(x+2 x_0)=e\sp{(\lambda_2-\lambda_1+\kappa\coth\kappa\nu_1) x}\,
{\sinh\kappa(\nu_1-x)\over\sinh\kappa \nu_1} \, \phi_1(2 x_0)
\label{sol1}
\end{equation}
where the exponential may be rewritten to yield
\begin{eqnarray*}
\lambda_2-\lambda_1+\kappa\coth\kappa\nu_1 &=&
{\phi_4\sp\prime(x_0)\over \phi_4(x_0)} -{N_2\over2}+\kappa\coth\kappa\nu_1=
{\phi_4\sp\prime(x_0)\over \phi_4(x_0)}+\kappa\coth(\kappa\nu_1/2)\\
&=&{\phi_5\sp\prime(x_0)\over \phi_5(x_0)} +{N_2\over2}+\kappa\coth\kappa\nu_1
={\phi_5\sp\prime(x_0)\over \phi_5(x_0)} +\kappa\tanh(\kappa\nu_1/2).
\end{eqnarray*}
We now have the information needed to determine $\phi_{4}(x)$ and
$\phi_{5}(x)$ via (\ref{eq:strategy}) which gives
\begin{eqnarray}
\phi_{4}(x+x_0)&=&{\sinh\kappa( \nu_1/2-x)\over \sinh(\kappa \nu_1/2)}
e\sp{\left(\phi_4\sp\prime(x_0)/ \phi_4(x_0)+\kappa\coth(\kappa\nu_1/2)
\right)x}\, \phi_{4}(x_0)
\label{sol2}\\
\phi_{5}(x+x_0)&=&{\cosh\kappa( \nu_1/2-x)\over \cosh(\kappa \nu_1/2)}
e\sp{\left(\phi_5\sp\prime(x_0)/ \phi_5(x_0)+\kappa\tanh(\kappa\nu_1/2)
\right)x}\, \phi_{5}(x_0)
\label{sol3}
\end{eqnarray}
Assembling this provides
\begin{lem}
The general analytic solution to (\ref{dexponential}) is given by
(\ref{sol1}), (\ref{sol2}) and (\ref{sol3}), where $x_0$ is a generic
point.
\end{lem}
\noindent{\bf Example 3.} We conclude by showing how our results determine
the solutions of the functional equation (\ref{eq:biggy}):
\[
\Psi _1(x+y)=\Psi _2(x+y) \phi_2(x)\phi_3(y) +\Psi_3(x+y) \phi_4(x)\phi_5(y).
\]
This equation encompasses as particular cases the equations
(\ref{dexponential}) (with $\Psi _2=\Psi _3=1$, $\phi _2(x)=\phi _4(x)$
and $\phi _3(x)=\phi _5(x)$) and (\ref{calogero}) (with
$(\phi _2(x)=\phi_3(x)$ and $\phi _4(x)=\phi _5(x)$) which have
already been discussed. Because of
this we will only consider the generic case
$\phi_2(x)\neq \lambda \phi_3(x)$, $\phi_4(x)\neq \gamma \phi_5(x)$
and $\Psi_2(x)\neq \delta \Psi_3(x)$
(where $\lambda ,\gamma ,\delta $ are constants) rather than these limits.
Our first step is to relate (\ref{eq:biggy}) to (\ref{functional}):
\begin{lem}
The functions $\Psi_m(x)$ ($m=1,2,3$) and $\phi_n(x)$ ($n=2,3,4,5$) give
a solution of equation (\ref{eq:biggy}) if and only if
\begin{equation}
\frac{\Psi_3(x+y)}{\Psi_2(x+y)}= -
{ { \biggl|\matrix{\phi_2(x)&\phi_2(y)\cr\phi_3(x)&\phi_3(y)\cr}\biggr|}
\over
{ \biggl|\matrix{\phi_4(x)&\phi_4(y)\cr\phi_5(x)&\phi_5(y)\cr}\biggr|} }
\label{eq:inter1}
\end{equation}
and
\begin{equation}
\frac{\Psi_1(x+y)}{\Psi _2(x+y)}=-
{ { \biggl|\matrix{\phi_2(x) \phi_5(x)&\phi_2(y)\phi_5(y)\cr
\phi_3(x) \phi_4(x)&\phi_3(y)\phi_4(y)\cr}\biggr|}
\over
{ \biggl|\matrix{\phi_4(x)&\phi_4(y)\cr\phi_5(x)&\phi_5(y)\cr}\biggr|} }
\label{eq:inter2}
\end{equation}
\end{lem}
\begin{proof} Assume first that the functions ${\Psi_m, \phi_n}$
give a solution of equation (\ref{eq:biggy}). Then after
interchanging $x$ and $y$ in (\ref{eq:biggy}) and subtracting
the result from (\ref{eq:biggy}) we obtain equation (\ref{eq:inter1}).
Upon substituting the formula for ${\Psi_3(x+y)}/{\Psi_2(x+y)}$
into (\ref{eq:biggy}) we arrive the formula (\ref{eq:inter2}).
In the other direction, let the functions ${\Psi_m, \phi_n}$
now satisfy (\ref{eq:inter1}) and (\ref{eq:inter2}). Upon writing the
right hand side of (\ref{eq:biggy}) as
\begin{equation}
\begin{array}{l}
\Psi_2(x+y)\phi_2(x)\phi_3(y)+\Psi_3(x+y)\phi_4(x)\phi_5(y)=\\
\quad\quad\quad\Psi_2(x+y)\left( \phi_2(x)\phi_3(y)
+\frac{\Psi_3(x+y)}{\Psi_2(x+y)}\phi_4(x)\phi_5(y)\right)
\end{array}
\label{eq:inter3}
\end{equation}
and using expression (\ref{eq:inter1}) for
${\Psi_3(x+y)}/{\Psi_2(x+y)}$ we find the term in brackets
in (\ref{eq:inter3}) rearranges to give precisely the
right hand side of (\ref{eq:inter2}); substituting for this
then yields (\ref{eq:biggy}) and therefore the required solution.
\end{proof}
We may now apply theorem 1 to show that if $\Psi_m(x)$
($m=1,2,3$) give a solution of (\ref{eq:biggy}) then we must have the
ratios
\begin{equation}
\frac{\Psi_1(x)}{\Psi_2(x)}= c_1e^{\lambda_1x}
\frac{\Phi(x;\mu_1)}{\Phi(x;\mu_2)},\quad
\frac{\Psi_3(x)}{\Psi_2(x)}=c_2e^{\lambda_2x}
\frac{\Phi(x;\mu_3)}{\Phi(x;\mu_4)}.
\label{eq:biggyratios}
\end{equation}
Further, because the denominators of (\ref{eq:inter1}) and (\ref{eq:inter2})
are the same, theorem 2 shows that $\mu_4=\mu_2$. Theorem 1 also
determines the functions $\phi_n(x)$ ($n=2,3,4,5$) up to a ${\cal G}$ action.
In fact, given three functions $\Psi_m(x)$
($m=1,2,3$) whose ratios satisfy (\ref{eq:biggyratios}) with $\mu_4=\mu_2$,
this is also sufficient to guarantee there are functions
$\phi_n(x)$ ($n=2,3,4,5$) for which (\ref{eq:biggy}) holds true.
To see this let us substitute these ratios into equation (\ref{eq:biggy})
to give
\begin{eqnarray}
c_1e^{\lambda_1 (x+y)}\Phi(x+y;\mu_1)=\Phi(x+y;\mu _2)
\phi_2(x)\phi_3(y)+\nonumber \\
c_2e^{\lambda_2 (x+y)}\Phi(x+y;\mu_3)\phi_4(x)\phi_5(y).
\label{eq:biggytr}
\end{eqnarray}
We will have established sufficiency
once we have shown how to construct the functions
$\phi_n(x)$.This will be achieved
utilizing various properties of the functions $\Phi (x;\nu )$.
\begin{lem}
The Baker-Akhiezer functions $\Phi (x;\nu )$ satisfy the equations
\begin{equation}
\Phi (x+\alpha ;\nu )=-e^{(\zeta(\alpha -\nu )+\zeta (\nu )-\zeta(\alpha ))x}
\Phi (\alpha ;\nu )\
\frac{ \Phi (x;\nu -\alpha )}{\Phi (-x;\alpha )}
\label{eq:biggypf2}
\end{equation}
and
\begin{eqnarray}
c e^{\gamma (x+y)}\,\Phi(x+y;\nu_1+\nu_2)=\Phi(x+y;\nu_1)\,\Phi(x;\nu_2)
\,\Phi (y;\nu_2)-\nonumber \\
\Phi (x+y;\nu_2)\,\Phi(x;\nu_1)\,\Phi(y;\nu_1),
\label{eq:biggylem}
\end{eqnarray}
where $c=\wp(\nu _2)-\wp(\nu _1)$ and
$\gamma =\zeta (\nu _1)+\zeta (\nu _2)-\zeta (\nu _1+\nu _2)$.
\end{lem}
These follow directly from the definition of $\Phi (x;\nu )$ and properties
of the Weierstrass sigma function; in particular
(\ref{eq:biggylem}) is a consequence of the \lq three term relation\rq\
of the sigma function (\cite[20.53, {\it Ex:5}]{WW}).
Upon setting $x\rightarrow x+\alpha$ in (\ref{eq:biggylem})
we obtain
\begin{equation}
\begin{array}{rl}
ce^{\gamma (x+y+\alpha)}\Phi (x+y+\alpha;\nu _1+\nu _2)&=\Phi(x+y+\alpha ;\nu_1)
\Phi (x+\alpha ;\nu_2)\Phi (y;\nu_2)\nonumber \\
&\quad-\Phi (x+y+\alpha ;\nu _2)\Phi (x+\alpha ;\nu _1)\Phi (y;\nu _1)
\label{eq:biggypf1}
\end{array}
\end{equation}
Now by substituting (\ref{eq:biggypf2}) in (\ref{eq:biggypf1})
and setting $\mu _1=\nu _1+\nu _2-\alpha$, $\mu_2=\nu_1-\alpha$ and
$\mu _3=\nu_2-\alpha$ one obtains after some rearrangement
\begin{equation}
\begin{array}{rl}
c\sp\prime e^{\lambda\sp\prime (x+y)}\Phi (x+y;\mu_1)=&
\Phi (x+y;\mu_2) \frac{\Phi (x;\mu_3)}{\Phi (-x;\mu_1-\mu_2-\mu_3)}
\Phi (y;\mu_1-\mu_2)+\nonumber\\
& c\sp{\prime\prime} e^{\lambda\sp{\prime\prime} (x+y)}
\Phi (x+y;\mu_3) \frac{\Phi (x;\mu_2)}{\Phi (-x;\mu_1-\mu_2-\mu_3)}
\Phi (y;\mu_1-\mu_3)
\label{eq:biggypf3}
\end{array}
\end{equation}
for appropriate constants $c\sp\prime, c\sp{\prime\prime},
\lambda\sp\prime, \lambda\sp{\prime\prime} $.
This is precisely of the desired form (\ref{eq:biggytr}).
Therefore we have shown
\begin{thm}
Given functions $\Psi_m(x)$ ($m=1,2,3$), there are functions $\phi_n(x)$
($n=2,3,4,5$) for which the functional equation
(\ref{eq:biggy}) is true if and only if the following ratios take place:
\begin{equation}
\frac{\Psi_1(x)}{\Psi_2(x)}= c_1e^{\lambda_1x}
\frac{\Phi(x;\mu_1)}{\Phi(x;\mu_2)},\quad
\frac{\Psi_3(x)}{\Psi_2(x)}=c_2e^{\lambda_2x}
\frac{\Phi(x;\mu_3)}{\Phi(x;\mu_2)},
\label{eq:biggythm}
\end{equation}
where $c_m,\lambda_m,$ ($m=1,2$) and $\mu_n$, ($n=1,2,3$) are free parameters.
\end{thm}
\vskip 1in
\noindent{\bf Acknowledgements:}
One of us (V.M.B.) thanks the Royal Society for a Kapitza Fellowship
in $1993$ and the EPSRC for a Visiting Fellowship in $1994-5$ over which time
this paper developed.
\newpage
|
\section{Introduction}
\la{Introduction}
An interesting special case of string/string duality
\cite{Luloop,Khurifour,Lublack,Minasian,Duffclassical,Khuristring,
Duffstrong} is
provided by the $D=10$ heterotic string compactified to $D=6$ on $T^4$
which is related by strong/weak coupling to the $D=10$ Type $IIA$
string compactified to $D=6$ on $K3$ \cite{Hull,Witten,Senssd,Harvey}.
The dilaton
${\tilde \Phi}$, metric ${\tilde G}_{MN}$ and $2$-form ${\tilde
B}_{MN}$ of the Type $IIA$ theory are related to those of the heterotic
theory, $\Phi$, $G_{MN}$ and $B_{MN}$, by
\cite{Luloop,Lublack,Minasian,Khuristring,Duffstrong}
\begin{eqnarray}
{\tilde \Phi}&=&-\Phi\nonumber\\
{\tilde G}_{MN}&=&e^{-\Phi}G_{MN}\nonumber\\
{\tilde H}&=&e^{-\Phi}*H\ ,
\la{dual}
\end{eqnarray}
where $M=0,\ldots,5$, $H=dB+\cdots$, ${\tilde H}=d{\tilde B}$ and $*$
denotes the Hodge dual. This ensures that the roles of $3$-form field
equations and Bianchi identities in one version of the corresponding
supergravity theory are interchanged in the other.
After further toroidal compactification to $D=4$ this automatically
accounts for the conjectured strong/weak coupling $SL(2,Z)$
duality in the resulting $D=4$, $N=4$ Type $IIA$ string and hence
for the $N=4$ Yang-Mills theories obtained by taking the global
limit \cite{Duffstrong}.
This is because $S$, the four-dimensional axion/dilaton
field, and $T$, the complex
Kahler form of the torus, are interchanged in going from the heterotic
to the Type $IIA$ theory. Moreover, while the
electric field strengths of the Kaluza-Klein gauge fields arising from
$G_{MN}$ are the same in both pictures, those of the ``winding" gauge fields
arising from $B_{MN}$ in the heterotic theory are replaced by their magnetic
duals in the Type $IIA$ theory. Thus the strong/weak coupling duality
of the Type $IIA$ string is just the target-space $SL(2,Z)_T$ of
the heterotic string.
However, the target space symmetry of the heterotic theory also
contains an $SL(2,Z)_U$ that acts on $U$, the complex structure of
the torus%
\footnote{In this paper, the phrase {\it U-duality} will be taken to
mean $SL(2,Z)_U$ called $SL(2,Z)_O$ in \cite{Duffstrong}. This
should not be confused with the $U$-duality of \cite{Hull} where it was
taken to mean the conjectured $E_7$ duality \cite{Luduality} of the
toroidally compactified Type $II$ string.}.
This suggests that, in addition to these $S$ and $T$ strings there
ought to be a third {\it $U$-string} whose axion/dilaton field is $U$
and whose strong/weak coupling duality is $SL(2,Z)_U$. From a
$D=6$ perspective, this seems strange since, instead of (\ref{dual}),
we now interchange $G_{45}$ and $B_{45}$. Moreover, of the two electric
field strengths which become magnetic, one is a winding gauge field and
the other is Kaluza-Klein! So such a duality has no $D=6$ Lorentz
invariant meaning. In fact, this $U$ string is a Type $IIB$ string, a
result which may also be understood from the point of view of
mirror%
\footnote{We are grateful to Xenia De La Ossa and Jan Luis for pointing
out that $T$--$U$ interchange is a mirror symmetry.}
symmetry: interchanging the roles of Kahler form and complex structure
(which is equivalent to inverting the radius of one of the two circles)
is a symmetry of the heterotic string but takes Type $IIA$ into Type
$IIB$ \cite{Dai,Dine}. In summary, if we denote the heterotic, $IIA$
and $IIB$ strings by $H,A,B$ respectively and the axion/dilaton,
complex Kahler form and complex structure by the triple $XYZ$ then we
have a triality between the $S$-string ($H_{STU}=H_{SUT}$), the
$T$-string ($B_{TUS}=A_{TSU}$) and the $U$-string ($A_{UST}=B_{UTS}$)
as illustrated in Fig. (1).
\begin{figure}
\centerline{\epsfbox{tri.eps}}
\medskip
\caption{String/string/string triality. The solid lines
correspond to string/string dualities and the dashed lines
represent mirror transformations.}
\end{figure}
The field theory limits of the heterotic string on $T^4$, the Type
$IIA$ string on $K3$ and the Type $IIB$ string on $K3$ are described by
certain $N=2,D=6$ supergravity theories described in section
(\ref{sec:N2D6}). As discussed in detail in section (\ref{sec:N4D4}),
each string in $D=4$ will then exhibit the same total symmetry
\begin{equation}
SL(2,Z)_S \times O(6,22;Z)_{TU} \supset
SL(2,Z)_S \times SL(2,Z)_T \times SL(2,Z)_U\ ,
\end{equation}
with the $28$ gauge field strengths and their duals
transforming as a $(2,28)$,
albeit with different interpretations for the three $SL(2,Z)$
factors. Note that there is a discrete symmetry under $T$--$U$
interchange, but there is no such $U$--$S$ or $S$--$T$ symmetry. As
discussed in \cite{Duffstrong}, it is the degrees of freedom associated
with going from $10$ to $6$ which are responsible for this lack of
$S$--$T$--$U$ democracy. This will also be reflected in the
Bogomol'nyi spectrum of electric and magnetic states that belong to the
short and intermediate $N=4$ supermultiplets. It is therefore
instructive to consider first the simpler situation where these modes
are truncated out. This we do first in section (\ref{sec:N1D6}) by
truncating the $N=2,D=6$ supergravities to $N=1,D=6$ and then in
section (\ref{sec:N2D4}) by reducing these supergravities to $D=4$. We
write down the action which describes the low energy limit of the
$S$-string; it exhibits an off-shell (perturbative) $SL(2,Z)_T
\times SL(2,Z)_U$ symmetry%
\footnote{The classical supergravities will in fact display continuous
symmetries such as $SL(2,R)$, but since these will be broken by quantum
corrections to discrete symmetries such as $SL(2,Z)$, we shall from
now on refer only to these.}
and an on-shell (non-perturbative) $SL(2,Z)_S$. Similarly, the
$T$-string action has an off-shell $SL(2,Z)_U \times SL(2,Z)_S$
and an on-shell $SL(2,Z)_T$, while the $U$-string action has an
off-shell $SL(2,Z)_S \times SL(2,Z)_T$ and an on-shell
$SL(2,Z)_U$. Aside from the pedagogical usefulness of this
$S$--$T$--$U$ symmetric truncation, which describes just $4$ of the
$28$ gauge fields, it will turn out that this theory and the resulting
$S$--$T$--$U$ symmetric Bogomol'nyi spectrum, discussed in section
(\ref{sec:N2Sol}), will find application in $N=2$ theories whose
Bogomol'nyi spectrum includes multiplets which were both short and
intermediate from the $N=4$ point of view. In particular we discuss the
extreme black hole spectrum \cite{Rahmfeld1,Senblacktorus,Kalloshpeet,Cvetic}.
In section (\ref{sec:N2Sol}) we provide a soliton interpretation of the
three strings. We identify the $S$-string with the {\it elementary
string} solution of \cite{Dabholkar}, the $T$-string with the {\it dual
solitonic string} solution of \cite{Khurifour} and the $U$-string with
(a limit of) the {\it stringy cosmic string} solution of \cite{Greene}.
In $D=3$ dimensions, all three strings are related by $O(4,4;Z)$
transformations.
In sections (\ref{sec:N2D6}), (\ref{sec:N4D4}), (\ref{sec:N4Bog}) and
(\ref{sec:N4Sol}) we repeat the exercise of sections (\ref{sec:N1D6}),
(\ref{sec:N2D4}), (\ref{sec:N2Bog}) and (\ref{sec:N2Sol}), now
including the full set of states. Section (\ref{sec:N2D6}) describes
the three $N=2$, $D=6$ supergravities: the actions in the heterotic and
Type $IIA$ cases (together with a duality dictionary relating the two
sets of fields) and the equations of motion in the case of Type $IIB$.
The compactification to $N=4$, $D=4$ of section (\ref{sec:N4D4})
reveals one or two surprises: although the $S$-string action has an
off-shell $O(6,22;Z)$ which continues to contain $SL(2,Z)_T
\times SL(2,Z)_U$, the $T$-string action has only an off-shell
$SL(2,Z)_U \times O(3,19;Z)$ which does not contain $SL(2,Z)_S$.
Similarly, the $U$-string action has only an $SL(2,Z)_T \times
O(3,19;Z)$ which does not contain $SL(2,Z)_S$. In short, none of
the actions is $SL(2,Z)_S$ invariant! This lack of off-shell
$SL(2,Z)_S$ in the Type $II$ actions can be traced to the presence
of the extra $24$ gauge fields which arise from the R-R sector of Type
$II$ strings: $S$-duality in the heterotic picture acts as an on-shell
electric/magnetic transformation on all $28$ gauge fields and continues
to be an on-shell transformation on the $24$ which remain unchanged
under the string/string/string triality%
\footnote{The absence of a $R\rightarrow 1/R$ $T$-duality symmetry of
the Type $II$ supergravity action in $D=9$ has been noted in
\cite{Bergshoeffduality}.}.
At first sight, this seems disastrous for deriving the strong/weak
coupling duality of the heterotic string from target space duality of
the Type $II$ string. The whole point was to explain a {\it
non-perturbative} symmetry of one string as a {\it perturbative}
symmetry of another \cite{Duffstrong}. Fortunately, all is not lost:
although $SL(2,Z)_S$ is not an off-shell symmetry of the Type $II$
supergravity
actions, it is still a symmetry of the Type $II$ string
theories. To see this we first note that $D=6$ general covariance is a
perturbative symmetry of the Type $IIB$ string and therefore that the
$D=4$ Type $IIB$ strings must have a perturbative $SL(2,Z)$ acting
on the complex structure of the compactifying torus. Secondly we note
that for both Type $IIB$ theories, $B_{TUS}$ and $B_{UTS}$, $S$ is the
complex structure field. Thus the $T$ string has $SL(2,Z)_U
\times SL(2,Z)_S$ and the $U$ string has
$SL(2,Z)_S \times SL(2,Z)_T$ as
required%
\footnote{We are grateful to Ashoke Sen for discussions on
these issues.}.
In this sense, four-dimensional string/string/string triality fills
a gap left by six-dimensional string/string duality: although duality
satisfactorily
explains the strong/weak coupling duality of the $D=4$ Type $IIA$
string in terms of the target space duality of the heterotic string,
the converse requires the Type $IIB$ ingredient.
Note that all of the three $SL(2,Z)_{(S,T,U)}$ take NS-NS states
into NS-NS states and that none can be identified with the conjectured
non-perturbative $SL(2,Z)_X$, where $X$ is the complex scalar of
the Type $IIB$ theory in $D=10$, which transforms NS-NS into R-R
\cite{Callan2,Hull,Witten}. However, this $SL(2,Z)_X$ is a
subgroup of $O(6,22;Z )$. Since this is a perturbative target space
symmetry of the heterotic string, the conjecture follows automatically
from the $D=4$ string/string/string triality hypothesis. Thus
we can say that evidence for this triality is evidence
not only for the electric/magnetic duality of all three $D=4$ strings
but also for the $SL(2,Z)_X$ of the $D=10$ Type $IIB$ string and
hence for {\it all} the conjectured non-perturbative symmetries of
string theory%
\footnote{One might object that in one case we have a
pre-compactification explanation but in the other only a
post-compactification explanation. However, having established
$SL(2,Z)_X$ in the compactified version, its presence in the
uncompactified version then follows by blowing up the extra dimensions
keeping fixed the complex $X$ field. We are grateful to Ashoke Sen
for this observation.}.
In section (\ref{sec:N4Bog}) we describe the $N=4,D=4$ Bogomol'nyi
spectrum. We generalize the heterotic string formula of Schwarz and
Sen, deriving the two $SL(2,Z)_S \times O(6,22;Z)_{TU}$
invariant central charges $Z_1$ and $Z_2$. This enables us to describe
the intermediate multiplets as well as the short ones, and once again we see
how the extreme black holes fit into this classification.
Section (\ref{sec:N4Sol}) generalizes (as far as is possible) the
soliton interpretation of section (\ref{sec:N2Sol}). But as discussed in
\cite{Duffstrong}, including the extra degrees of freedom in going from
$10$ to $4$ causes problems in identifying the soliton zero modes.
Although it is straightforward to find the heterotic string as a
soliton of Type $II$, the converse is more problematical
\cite{Senssd,Harvey}. In three dimensions, the $O(4,4;Z)$
generalizes to $O(8,24;Z)$
\cite{Luduality,Marcus,Rahmfeld1,Sen7,Rahmfeld2}.
Four-dimensional {\it string/string/string triality} was announced by
one of us (MJD) at the PASCOS 95 conference in Baltimore and at the
SUSY 95 conference in Paris \cite{Duffelectric}. Related results have
been obtained independently by Aspinwall and Morrison
\cite{Aspinwall2}.
\section{$N=1$ supergravity in $D=6$}
\label{sec:N1D6}
As a good guide to the kind of dualities one might expect in string
theory, it pays to look first at the corresponding supergravity
theories. We therefore review some properties of $D=6$ supergravity
\cite{Salam}. The theories of interest, which follow
either from $T^4$ compactification of the $D=10$ heterotic string or
$K3$ compactification of Type $II$, will be $N=2$ supergravities in
$D=6$ which yields $N=4$ in $D=4$. All these theories are non-minimal in
the sense that they contain additional $N=2$ gauge or matter multiplets.
Since such additional matter destroys the $S$--$T$--$U$ symmetry of the
four-dimensional string we begin by examining an $N=1$ subset common to all
the models of interest. We return to the full $N=2$ theory in section
(\ref{sec:N2D6}).
In terms of six-dimensional $N=1$ representations, we focus on the
supergravity multiplet $(G_{MN},\Psi^{+A}{}_M,B^+{}_{MN})$ and the
self-dual tensor multiplet $(B^-{}_{MN},\chi^{+A},\Phi)$. The index
$A=1,2$ labels the $\bf 2$ of $Sp(2)$ and both spinors are symplectic
Majorana-Weyl. The $2$-forms $B^+{}_{MN}$ and $B^-{}_{MN}$ have
$3$-form field strengths that are self-dual or anti-self-dual,
respectively. Only with the combination of one supergravity multiplet
and one self-dual tensor multiplet do we have a conventional Lagrangian
formulation. In this case the bosonic fields correspond to the
graviton, antisymmetric tensor and dilaton of string theory. This
simpler theory will not only serve as a warm-up exercise for
understanding the $N=4,D=4$ superstrings but is interesting in its own
right for understanding the $N=2,D=4$ strings.
There are three theories to consider, each with the same number of
physical degrees of freedom. The first two theories arise from the
truncation of the non-chiral $N=2$ supergravity and are related by
duality: the first has the usual $3$-form field strength $H$ and the
second has the dual field strength ${\tilde H}=e^{-\Phi}*H$. The third
theory comes from the truncation of the chiral $N=2$ supergravity.
While the full chiral $N=2$ theory does not admit a covariant
Lagrangian, the $N=1$ truncation, involving the combination of the
supergravity and tensor multiplet given above, may be written in a
conventional form. In anticipation of their future application, we
shall call these theories $H$, $A$ and $B$, respectively.
Denoting the $D=6$ spacetime indices by
$(M,N=0,...,5)$, the bosonic part of the usual action takes the
form
\begin{equation}
I_H=\frac{1}{2\kappa^2}\int d^6x \sqrt{-G}e^{-\Phi}\left[
R_G+G^{MN}\partial_M\Phi\partial_N\Phi
-\frac{1}{12}G^{MQ}G^{NR}G^{PS}H_{MNP}H_{QRS}\right]\ .
\la{H}
\end{equation}
$H$ is the curl of the $2$-form $B$
\begin{equation}
H=dB
\la{b}
\end{equation}
(at this point there is no Chern-Simons correction). The metric
$G_{MN}$ is related to the canonical Einstein metric $G^c{}_{MN}$ by
\begin{equation}
G_{MN}=e^{\Phi/2}G^c{}_{MN}\ .
\la{metric}
\end{equation}
Similarly, the dual supergravity action is given by
\begin{equation}
I_A=\frac{1}{2\kappa^2}\int d^6x \sqrt {-\tilde G}e^{-\tilde\Phi}\left[
R_{\tilde G}+\tilde G^{MN}\partial_M\tilde{\Phi}\partial_N\tilde{\Phi}
-\frac{1}{12}\tilde G^{MQ}\tilde G^{NR}\tilde G^{PS}
\tilde H_{MNP}\tilde H_{QRS}\right]\ .
\la{A}
\end{equation}
$\tilde H$ is also the curl of a 2-form $\tilde B$
\begin{equation}
\tilde H=d\tilde B\ .
\la{d}
\end{equation}
The dual metric $\tilde G_{MN}$ is related to the canonical Einstein
metric by
\begin{equation}
\tilde G_{MN}=e^{\tilde\Phi/2}G^c{}_{MN}\ .
\la{e}
\end{equation}
The two supergravities are related by:
\begin{eqnarray}
\tilde\Phi&=&-\Phi\nonumber\\
\tilde G_{MN}&=&e^{-\Phi}G_{MN}\nonumber\\
\tilde H&=&e^{-\Phi}\, {\ast H}\ ,
\la{duality}
\end{eqnarray}
where $\ast$ denotes the Hodge dual. (Since the last equation is
conformally invariant, it is not necessary to specify which metric is
chosen in forming the dual.) This ensures that the roles of field
equations and Bianchi identities in the one version of supergravity are
interchanged in the other. The combined field equations and Bianchi
identities therefore exhibit a discrete symmetry under interchange of
$\Phi\rightarrow -\Phi$, $G\to \tilde G$ and $H\rightarrow \tilde H$.
Finally, while the third theory is unrelated to the other two (at least
in $D=6$), at this level of truncation it has a bosonic action with a
form similar to that of $I_A$. One subtlety is worth mentioning, however.
Since this model arises from a truncation of the compactified Type $IIB$
string which has a {\it complex} 3-form field strength in ten dimensions,
there is some ambiguity in the identification of the dilaton
$\skew0\tilde{\tilde\Phi}$ and 3-form $\skew4\tilde{\tilde H}$ of model B, given in
the action
\begin{equation}
I_B=\frac{1}{2\kappa^2}\int d^6x
\sqrt{-\skew4\tilde{\tilde G}} e^{-\skew0\tilde{\tilde\Phi}}\left[R_{\skew4\tilde{\tilde G}}
+\skew4\tilde{\tilde G}^{MN}\partial_M\skew0\tilde{\tilde\Phi}\partial_N\skew0\tilde{\tilde\Phi}
-\frac{1}{12}\skew4\tilde{\tilde G}^{MQ}\skew4\tilde{\tilde G}^{NR}\skew4\tilde{\tilde G}^{PS} \skew4\tilde{\tilde H}_{MNP}\skew4\tilde{\tilde H}_{QRS}\right]\ .
\la{eq:B}
\end{equation}
In particular, the $SL(2,Z)_X$ symmetry of the Type $IIB$
supergravity will mix $\skew4\tilde{\tilde H}$ with its counterpart. Nevertheless, from
a stringy viewpoint, we may identify $e^{\skew0\tilde{\tilde\Phi}}$ as the string loop
expansion parameter and $\skew4\tilde{\tilde H}$ as the 3-form field strength arising
from the NS-NS sector of the string. This provides a unique definition
of the truncated action, (\ref{eq:B}). Note that there is no $D=6$
Lorentz invariant dictionary between the fields $(\skew0\tilde{\tilde\Phi}, \skew4\tilde{\tilde G},\skew4\tilde{\tilde H})$
and $(\Phi,G,H)$ or $(\tilde\Phi, \tilde G,\tilde H)$.
\section{The $S$-$U$-$T$ symmetric theory in $D=4$}
\label{sec:N2D4}
Now let us first consider the H theory, dimensionally
reduced to $D=4$. The combination of the six-dimensional $N=1$ supergravity
and tensor multiplets reduce to give the $D=4$, $N=2$ graviton multiplet
with helicities $(\pm2,2(\pm{3\over2}),\pm1)$ and three vector multiplets
with helicities $(\pm1,2(\pm{1\over2}),2(0))$. In order to make this
explicit, we use a standard decomposition of the six-dimensional
metric
\begin{equation}
G_{MN}=\pmatrix{g_{\mu\nu}+A_\mu^mA_\nu^nG_{mn}&A_\mu^mG_{mn}\cr
A_\nu^nG_{mn}&G_{mn}}\ ,
\label{metricred}
\end{equation}
where the spacetime indices are ${\mu,\nu}=0,1,2,3$ and the internal indices
are $m,n=1,2$. The remaining two vectors arise from the reduced $B$ field
\begin{equation}
B_{MN}=\pmatrix{B_{\mu\nu}+{\textstyle{1\over2}}(A_\mu^mB_{m\nu}+B_{\mu n}A_\nu^n)&
B_{\mu n}+A_\mu^mB_{mn}\cr
B_{m\nu}+B_{mn}A_\nu^n&B_{mn}}\ .
\end{equation}
Four of the six resulting scalars are moduli of the 2-torus. We parametrize
the internal metric and $2$-form as
\begin{equation}
G_{mn}=e^{\rho-\sigma}\left(
\begin{array}{cc}
e^{-2\rho}+c^2&-c\\
-c&1
\end{array}\right)\ ,
\end{equation}
and
\begin{equation}
B_{mn}=b\,\epsilon_{mn}\ .
\end{equation}
The four-dimensional metric, given by $g_{\mu\nu}$, is
related to the four-dimensional canonical Einstein, $g^c_{\mu\nu}$, metric by
$g_{\mu\nu}=e^{\eta}g^c{}_{\mu\nu}$ where $\eta$ is the four-dimensional
shifted dilaton:
\begin{equation}
e^{-\eta}=e^{-\Phi}\sqrt{\det G_{mn}}=e^{-(\Phi+\sigma)}\ .
\end{equation}
Thus the remaining two scalars are the dilaton $\eta$ and axion $a$ where
the axion field $a$ is defined by
\begin{equation}
\epsilon^{\mu\nu\rho\sigma}\partial_{\sigma}a=
\sqrt{-g}e^{-\eta}g^{\mu\sigma}g^{\nu\lambda}g^{\rho\tau}
H_{\sigma\lambda\tau}\ ,
\label{eq:Haxion}
\end{equation}
where
\begin{eqnarray}
H_{\sigma\lambda\tau}&=&3(\partial_{[\sigma}B_{\lambda\tau]}+\frac{1}{2}
A_{[\sigma}^m F_{\lambda\tau] m}+\frac{1}{2}B_{m [\sigma} F_{\lambda\tau]}^m)
\nonumber} \def\bd{\begin{document}} \def\ed{\end{document} \\
F_{\lambda\tau}^m&=&\partial_\lambda A_\tau^m-\partial_\tau A_\lambda^m \\
F_{\lambda\tau m}&=&\partial_\lambda B_{m\tau}-\partial_\tau B_{m\lambda}\
. \nonumber} \def\bd{\begin{document}} \def\ed{\end{document}
\end{eqnarray}
[...] denotes antisymmetrization with weight one.
We may now combine the above six scalars into the complex axion/dilaton
field $S$, the complex Kahler form field $T$ and the complex structure
field $U$ according to
\begin{eqnarray}
S&=&S_1+iS_2=a+ie^{-\eta}\nonumber\\
T&=&T_1+iT_2=b+ie^{-\sigma}\nonumber\\
U&=&U_1+iU_2=c+ie^{-\rho}\ .
\end{eqnarray}
This complex parametrization allows for a natural transformation under the
various $SL(2,Z)$ symmetries. The action of $SL(2,Z)_S$ is given by
\begin{equation}
S \rightarrow \frac{aS+b}{cS+d}\ ,
\la{sl2zs}
\end{equation}
where $a,b,c,d$ are integers satisfying $ad-bc=1$, with similar
expressions for $SL(2,Z)_T$ and $SL(2,Z)_U$. Defining the
matrices ${\cal M}_S$, ${\cal M}_T$ and ${\cal M}_U$ via
\begin{equation}
{\cal M}_S=\frac{1}{S_2}
\left(
\begin{array}{cc}
1 & S_1\\
S_1 & |S|^2
\end{array}
\right)\ ,
\label{eq:sl2mat}
\end{equation}
the action of $SL(2,Z)_S$ now takes the form
\begin{equation}
{\cal M}_S\rightarrow \omega_S{}^T{\cal M}_S\omega_S\ ,
\end{equation}
where
\begin{equation}
\omega_S=
\left(
\begin{array}{cc}
d& b\\
c & a
\end{array}
\right)\ ,
\end{equation}
with similar expressions for ${\cal M}_T$ and ${\cal M}_U$.
We also define the $SL(2,Z)$ invariant tensors
\begin{equation}
\epsilon_S=\epsilon_T=\epsilon_U=
\left(
\begin{array}{cc}
0& 1\\
-1 & 0
\end{array}
\right)\ .
\end{equation}
The fundamental supergravity (\ref{H}) now becomes
\begin{eqnarray}
I_{STU}&=&\frac{1}{16\pi G}\int d^4x\sqrt{-g}e^{-\eta}\Bigl[
R_g + g^{\mu\nu}\partial_{\mu}\eta\partial_{\nu}\eta
-\frac{1}{12}g^{\mu\lambda}g^{\nu\tau}g^{\rho\sigma}
H_{\mu\nu\rho}H_{\lambda\tau\sigma}\nonumber\\
&&\kern9.3em
+\frac{1}{4}{\rm Tr} (\partial{\cal M}_T{}^{-1}\partial {\cal M}_T)
+\frac{1}{4}{\rm Tr} (\partial{\cal M}_U{}^{-1}\partial {\cal M}_U)\nonumber\\
&&\kern9.3em
-\frac{1}{4}{F_S}_{\mu\nu}{}^T({\cal M}_T \times {\cal M}_U){F_S}^{\mu\nu}
\Bigr]\ .
\la{S}
\end{eqnarray}
\begin{figure}
\epsfxsize=4in
\centerline{\epsfbox{cube.eps}}
\medskip
\caption{The cube of triality. All field strengths are given in $S$-variables.}
\end{figure}
The four $U(1)$ gauge fields $A_S^a$ are given by
${A_S^1}_\mu=B_{4\mu}, \, {A_S^2}_\mu=B_{5\mu}, \, {A_S^3}_\mu=A_\mu^ 5,
\, {A_S^4}_\mu=-A_\mu^4$. The three-form becomes
$H_{\mu\nu\rho}=3(\partial_{[\mu}B_{\nu\rho]}
-{\textstyle{1\over2}} A_{S[\mu}{}^T (\epsilon_T\times\epsilon_U){F_S}_{\nu\rho]})$.
This action is
manifestly invariant under $T$-duality and $U$-duality, with
\begin{equation}
{F_S}_{\mu\nu}\rightarrow
(\omega_T{}^{-1}\times\omega_U{}^{-1}){F_S}_{\mu\nu}\ , \, \qquad
{\cal M}_{T/U}\rightarrow \omega_{T/U}^T \, {\cal M}_{T/U} \, \omega_{T/U}\ ,
\end{equation}
and with $\eta$, $g_{\mu\nu}$ and $B_{\mu\nu}$ inert. Its equations of
motion and Bianchi identities (but not the action itself) are also
invariant under $S$-duality, with $T$ and $g^c{}_{\mu\nu}$ inert and
with
\begin{equation}
\left(
\begin{array}{c}
{{F_S}}_{\mu\nu}{}^a\\
{\widetilde{F}_S}{}_{\mu\nu}{}^a
\end{array}
\right)
\rightarrow \omega_S^{-1}
\left(
\begin{array}{c}
{{F_S}}_{\mu\nu}{}^a\\
{\widetilde{F}_S}{}_{\mu\nu}{}^a
\end{array}
\right)\ ,
\end{equation}
where
\begin{equation}
{\widetilde{F}_S}{}_{\mu\nu}{}^{a}=-S_2[({\cal M}_T{}^{-1} \times {\cal
M}_U{}^{-1})(\epsilon_T \times \epsilon_U)]^a{}_b *
{F_S}_{\mu\nu}{}^{b}-S_1 {F_S}_{\mu\nu}{}^{a} \ .
\end{equation}
Thus $T$-duality transforms Kaluza-Klein electric charges
$({F_S}^3,{F_S}^4)$ into winding electric charges $({F_S}^1,{F_S}^2)$
(and Kaluza-Klein magnetic charges into winding magnetic charges),
$U$-duality transforms the Kaluza-Klein and winding electric charge of
one circle $({F_S}^3,{F_S}^2)$ into those of the other
$({F_S}^4,{F_S}^1)$ (and similarly for the magnetic charges) but
$S$-duality transforms Kaluza-Klein electric charge $({F_S}^3,{F_S}^4)$
into winding magnetic charge $({\tilde {F_S}}^3,{\tilde {F_S}}^4)$ (and
winding electric charge into Kaluza-Klein magnetic charge). In summary
we have $SL(2,Z)_T \times SL(2,Z)_U$ and $T \leftrightarrow U$
off-shell but $SL(2,Z)_S \times SL(2,Z)_T \times SL(2,Z)_U$
and an $S$--$T$--$U$ interchange on-shell. The $S \leftrightarrow T$
part arises from the discrete on-shell symmetry $\Phi\rightarrow -\Phi$,
$G\to \tilde G$ and $H\rightarrow \tilde H$ in $D=6$.
Now consider the two actions obtained by cyclic permutation of the fields
$S,T,U$:
\begin{eqnarray}
I_{TUS}&=&\frac{1}{16\pi G}\int d^4x\sqrt{-{\tilde g}}e^{-\sigma}\Bigl[
R_{\tilde g} + {\tilde g}^{\mu\nu}\partial_{\mu}\sigma\partial_{\nu}\sigma
-\frac{1}{12}
{\tilde g}^{\mu\lambda}{\tilde g}^{\nu\tau}{\tilde g}^{\rho\sigma}
{\tilde H}_{\mu\nu\rho}{\tilde H}_{\lambda\tau\sigma}\nonumber\\
&&\kern9.3em
+\frac{1}{4}{\rm Tr} (\partial{\cal M}_U{}^{-1}\partial {\cal M}_U)
+\frac{1}{4}{\rm Tr} (\partial{\cal M}_S{}^{-1}\partial {\cal M}_S)\nonumber\\
&&\kern9.3em
-\frac{1}{4}{F_T}_{\mu\nu}{}^T({\cal M}_U \times {\cal M}_S){F_T}^{\mu\nu}
\Bigr]\ ,
\la{T}
\end{eqnarray}
and
\begin{eqnarray}
I_{UST}&=&\frac{1}{16\pi G}\int d^4x\sqrt{-\skew3\tilde{\tilde g}}
e^{-\rho}\Bigl[R_{\skew3\tilde{\tilde g}}+\skew3\tilde{\tilde g}^{\mu\nu}\partial_\mu\rho\partial_\nu\rho
-\frac{1}{12}\skew3\tilde{\tilde g}^{\mu\lambda} \skew3\tilde{\tilde g}^{\nu\tau} \skew3\tilde{\tilde g}^{\rho\sigma}
\skew4\tilde{\tilde H}_{\mu\nu\rho}\skew4\tilde{\tilde H}_{\lambda\tau\sigma}
\nonumber\\
&&\kern9.3em
+\frac{1}{4}{\rm Tr} (\partial{\cal M}_S{}^{-1}\partial {\cal M}_S)
+\frac{1}{4}{\rm Tr} (\partial{\cal M}_T{}^{-1}\partial {\cal M}_T)\nonumber\\
&&\kern9.3em
-\frac{1}{4}{F_U}_{\mu\nu}{}^T({\cal M}_S \times {\cal M}_T){F_U}^{\mu\nu}
\Bigr]\ .
\la{U}
\end{eqnarray}
The $D=6$ interpretation of these actions is as follows. The action
$I_{TSU}=I_{TUS}$ is obtained by reducing the dual $A$ theory
(\ref{A}), where the four dimensional dual metric is given by ${\tilde
g}_{\mu\nu}=e^{\sigma}g^c{}_{\mu\nu}$ and the $3$-form field strength
${\tilde H}$ is related to the pseudoscalar field $b$ by
\begin{equation}
\epsilon^{\mu\nu\rho\sigma}\partial_{\sigma}b=
\sqrt{-\tilde g}e^{-\sigma}\tilde g^{\mu\sigma}\tilde g^{\nu\lambda}
\tilde g^{\rho\tau} \tilde H_{\sigma\lambda\tau}\ .
\end{equation}
However, since mirror symmetry interchanges $A$ and $B$ it also yields
the field equations obtained by reducing the field equations of the $B$
theory but with $S$ and $U$ interchanged. Similarly, the action
$I_{UST}=I_{UTS}$ yields the field equations obtained by reducing the
$B$ theory, where the four dimensional metric is now given by
$\skew3\tilde{\tilde g}_{\mu\nu} = e^{\rho}g^c{}_{\mu\nu}$ and the $3$-form field strength
$\skew4\tilde{\tilde H}$ is related to the pseudoscalar field $c$ by
\begin{equation}
\epsilon^{\mu\nu\rho\sigma}\partial_{\sigma}c= \sqrt{-\skew3\tilde{\tilde g}}
e^{-\rho}{\skew3\tilde{\tilde g}}^{\mu\sigma}{\skew3\tilde{\tilde g}}^{\nu\lambda}
{\skew3\tilde{\tilde g}}^{\rho\tau} {\skew4\tilde{\tilde H}}_{\sigma\lambda\tau}\ .
\end{equation}
Once again, however, by mirror symmetry this is equivalent to reducing
the $A$ theory with $S$ and $T$ interchanged. The
relation between the field strengths ${F_S}$, ${F_T}$ and ${F_U}$ is
given in Table 1 and Figure 2. Figure 2 visualizes the
connection of all three strings. Each side of the cube
corresponds to electric or
magnetic $S$, $T$ or $U$ strings. Each dimension is related to one duality.
To get from one side to an adjacent one, two fields need to be dualized.
Mirror symmetry takes the cube into its mirror.
\begin{table}
$$
\begin{array}{ccccccc}
\rm axion/&\rm Kahler&\rm complex&\multispan4\hfil\rm gauge\ fields\hfil\\
\rm dilaton&\rm form&\rm structure&&&&\\
&&&&&&\\
S&T&U&{F_S}^3&-{F_S}^4&{F_S}^1&-{F_S}^2\\
S&U&T&{F_S}^3&F_S^1&-{F_S}^4&-{F_S}^2\\
U&S&T&{F_S}^3&{F_S}^1&-{\tilde {F_S}}^3&-\tilde{F_S}^1\\
U&T&S&{F_S}^3&-\tilde{F_S}^3&{F_S}^1&-\tilde{F_S}^1\\
T&U&S&{F_S}^3&-\tilde{F}_S^3&-F_S^4&\tilde{F}_S^4\\
T&S&U&{F_S}^3&-F_S^4&-\tilde{F}_S^3&{\tilde{F}_S}^4
\end{array}
$$
\label{table1}
\caption{Triality}
\end{table}
\section{The Bogomol'nyi Spectrum}
\label{sec:N2Bog}
It is now straightforward to write down an $S$--$U$--$T$ symmetric
Bogomol'nyi mass formula. Let us define electric and magnetic charge
vectors $\alpha_S^a$ and $\beta_S^a$ associated with the field strengths
${{F_S}}^a$ and ${\tilde {F_S}}^a$ in the standard way.
The electric and magnetic charges $Q_S^a$ and $P_S^a$ are
given by
\begin{equation} {F_{S}}_{0r}^a\sim\frac{Q_S^a}{r^2} \, \qquad
*{F_{S}}_{0r}^a\sim\frac{P_S^a}{r^2}\ ,
\end{equation}
giving rise to the charge vectors
\begin{equation}
\pmatrix{\a_S^a\cr \b_S^a}=\pmatrix{ S_2^{(0)} {\cal M}_T^{-1}
\times {\cal M}_U^{-1} & S_1^{(0)} \e_T \times \e_U \cr 0 &
-\e_T \times \e_U }^{ab} \pmatrix{Q_S^b \cr P_S^b}.
\end{equation}
For our purpose it is useful to define a generalized charge vector
$\gamma^{a{\tilde a}
{\tilde{\tilde a}}}$ via
\begin{equation}
\left(
\begin{array}{c}
\gamma^{111}\\
\gamma^{112}\\
\gamma^{121}\\
\gamma^{122}\\
\gamma^{211}\\
\gamma^{212}\\
\gamma^{221}\\
\gamma^{222}
\end{array}
\right)
=
\left(
\begin{array}{c}
-\beta_S^1\\
-\beta_S^2\\
-\beta_S^3\\
-\beta_S^4\\
\alpha_S^1\\
\alpha_S^2\\
\alpha_S^3\\
\alpha_S^4
\end{array}
\right)\ ,
\end{equation}
transforming as
\begin{equation}
\gamma^{a{\tilde a}{\tilde {\tilde a}}}\rightarrow
\omega_S{}^a{}_b
\omega_T{}^{\tilde a}{}_{\tilde b}
\omega_U{}^{\tilde {\tilde a}}{}_{\tilde {\tilde b}}
\gamma^{b{\tilde b}{\tilde {\tilde b}}}\ .
\end{equation}
Then the mass formula is
\begin{equation}
m^2=\frac{1}{16}\gamma^T({\cal M}_S{}^{-1}{\cal M}_T{}^{-1}{\cal M}_U{}^{-1}
-{\cal M}_S{}^{-1}{\epsilon}_T{\epsilon}_U
-{\epsilon}_S{\cal M}_T{}^{-1}{\epsilon}_U
-{\epsilon}_S{\epsilon}_T{\cal M}_U{}^{-1})\gamma\ .
\la{us}
\end{equation}
Although all three theories have the same mass spectrum, there is
clearly a difference of interpretation with electrically charged
elementary states in one picture being solitonic monopole or dyon
states in the other. This agrees with the $N=2$ Bogomol'nyi formula of
Ceresole {\it et al.}\ \cite{Ceresole} and is a truncation of the generalized
$N=4$ mass formula derived from first principles
in section (\ref{sec:N4Bog}). Note, however, that this is {\it not}
a truncation of the $N=4$ Bogomol'nyi formula of Schwarz and Sen
\cite{Schwarz1,Sen5}. In particular,
we note that
although both formulas have $SL(2,Z)_S \times SL(2,Z)_T \times
SL(2,Z)_U$, even the truncated
Schwarz-Sen formula (\ref{schwsen}) is only symmetric under $T$--$U$
interchange and not $S$--$T$--$U$. To understand this, we recall
that in $N=4$ supersymmetry, we have two central charges $Z_1$ and
$Z_2$. There are three kinds of massive multiplets: short,
intermediate and long according as $(m=|Z_1|=|Z_2|)$, $(m=|Z_1|>|Z_2|)$
or $(m>|Z_1|,|Z_2|)$. The Schwarz-Sen formula refers only to the short
multiplets. In $N=2$, however, we have only one central charge $Z$.
There are only short and long multiplets according as $m=|Z|$ or
$m>|Z|$. States that were only intermediate in the $N=4$ theory may
thus become short in the truncation to $N=2$.
A nice example of this
phenomenon is provided by the extreme Reissner-Nordstrom black hole
(dilaton coupling $a=0$)
which in string theory is dyonic with charge vectors
$\alpha=(1,0,0,-1)$ and $\beta=(0,-1,-1,0)$ \cite{Rahmfeld1}. It
belongs to an intermediate multiplet in the $N=4$ theory and is
therefore absent from the Sen-Schwarz spectrum but belongs to a short
multiplet in the $N=2$ theory and appears in the spectrum (\ref{us}).
The two $N=4$ central charges are given in section (\ref{sec:N4Bog}).
Since we have identified the Reissner-Nordstrom black hole in the
$N=2$ spectrum, it is natural to ask which other black holes satisfy
(\ref{us}). Besides $a=0$, the supersymmetric
dilaton coupling parameters are $a=\sqrt{3},1,1/\sqrt{3}$
\cite{Khurinew,Rahmfeld1,Horowitz2,Khuriscatter}. It turns out that all of the
corresponding states indeed satisfy the Bogomol'nyi bound and therefore
preserve 1/2 of the supersymmetries in the $N=2$ theory.
The $a=\sqrt{3}$ black hole has charge vectors $\a=(1,0,0,0)$,
$\b=(0,0,0,0)$. To cut a long story
short we set all the VEV's to zero and find its mass to be (in our units)
$m=1/4$, according to
\begin{equation}
m^2=\frac{Q^2}{4(1+a^2)}\ ,
\end{equation}
where $Q$ is the charge of the effective field strength.
Mass and charges are obviously related by (\ref{us}).
The mass of the
electrically charged $a=1$ black hole with $\a=(1,0,0,-1)$ is
$m=1/2$ \cite{Rahmfeld1}
which agrees also with (\ref{us}). Like the $a=\sqrt{3}$ black hole, this
solution
is elementary for the S-string, but it is dyonic for the $T$- and $U$-strings.
Further dynamical
evidence for the identification of $a=\sqrt{3}$ and $a=1$ black holes
with elementary heterotic
$N_L=1$ and $N_L>1$ string states \cite{Rahmfeld1} has recently
been given in \cite{Khuriscatter}.
Finally, the $a=1/\sqrt{3}$ black hole is dyonic in all pictures.
Its charge vectors are $\a=(1,0,0,-1)$ and $\b=(0,-1,0,0)$. The mass
is $m=3/4$ which can be verified by truncating the supergravity theory to
one effective field strength $\sqrt{3}F=F_S^1=-F_S^4=\tilde{F}_S^2$ along the
lines of \cite{Rahmfeld1}. A quick comparison with the Bogomol'nyi formula
proves that the $a=1/\sqrt{3}$ black hole preserves indeed 1/2
of the supersymmetries in $N=2$. As described in \cite{Rahmfeld1}, the
mass and charge assignments of the $a=\sqrt{3},1,1/\sqrt{3}$ and $0$
black holes are compatible with their interpretations as $1,2,3$ and
$4$-particle bound states with zero binding energy.
\section{Soliton Interpretation}
\label{sec:N2Sol}
Four-dimensional string/string/string/triality suggests that it ought
to be possible to describe the $S$-string, $T$-string and $U$-string as
elementary and solitonic solutions directly in four dimensions. This is
indeed the case. The $H$ action (\ref{H}) admits as an elementary
solution the $S$-string
\begin{eqnarray}
ds^2&=&e^{\eta}(-d\tau^2+d\sigma^2)+dzd\bar{z}\nonumber\\
S&=&a+ie^{-\eta}=\frac{1}{2\pi i}\ln\frac{r}{r_0}\ ,
\la{string1}
\end{eqnarray}
where $z=x_2+ix_3$ corresponds to the transverse directions and
$r=|z|$. It also admits as a soliton solution the dual $T$-string
\begin{eqnarray}
ds^2&=&-d\tau^2+d\sigma^2+e^{-\sigma}dzd\bar{z}\nonumber\\
T&=&b+ie^{-\sigma}=\frac{1}{2\pi i}\ln\frac{r}{r_0}\ .
\la{string2}
\end{eqnarray}
Furthermore, it admits as a soliton solution the $U$-string
\begin{eqnarray}
ds^2&=&-d\tau^2+d\sigma^2+e^{-\rho}dzd\bar{z}\nonumber\\
U&=&c+ie^{-\rho}=\frac{1}{2\pi i}\ln\frac{r}{r_0}\ .
\la{string3}
\end{eqnarray}
We recognize the $S$-string as the {\it elementary string} solution of
\cite{Dabholkar} and the $T$-string as the {\it dual string} solution
of \cite{Khurifour} but the $U$-string is given by a limit of the {\it
stringy cosmic string} of \cite{Greene} where the fields $\rho$ and $c$
are simply given by the internal metric
\begin{equation}
\sqrt{G}G^{-1}={\cal M}_U=e^{\rho}
\left(
\begin{array}{cc}
1 & c\\
c & c^2+e^{-2\rho}
\end{array}
\right)\ .
\end{equation}
Consequently, the $U$-string is a solution of pure gravity in $D=6$ as
discussed in \cite{Greene}.
It follows that the $A$ action (\ref{A}) admits the $T$-string as
the elementary solution and the $S$- and $U$-strings as the solitonic
solutions and that the $B$ action (\ref{eq:B}) admits the $U$-string as
the elementary solution and the $T$- and $S$-strings as the solitonic
solutions. Note that we may generate new $S$, $T$- and $U$-string
solutions by making $SL(2,Z)_S$ transformations on (\ref{string1}),
$SL(2,Z)_{T}$ transformations on (\ref{string2}) and
$SL(2,Z)_{U}$ transformations on (\ref{string3}). So there is
really an $SL(2,Z)$ family of solutions for each string. Once again,
all this is consistent with string/string/string triality.
The fundamental string solution given in (\ref{string1}) corresponds to
the case where all four gauge fields $(F_S{}^1,F_S{}^2,F_S{}^3,F_S{}^4)$
have been set to zero. But as described in \cite{Sen2} a more general
solution with non-vanishing gauge fields may be generated by making
$O(3,3)$ transformations on the neutral solution. Such deformations
are possible since the original solution is independent of $x^0$ as
well as $x^4$ and $x^5$. However, since we want to keep the asymptotic
values of the field configurations fixed, this leaves us with an
$O(2,1) \times O(2,1)$ subgroup. Not every element of this subgroup
generates a new solution; there is an $O(2) \times O(2)$ subgroup
that leaves the solution invariant. Thus the number of independent
deformations is given by the dimension of the coset space $O(2,1)
\times O(2,1)/O(2) \times O(2)$ which is equal to four, corresponding
to the four electric charges of $U(1)^4$. Exactly analogous statements
now apply to the $T$-string (\ref{string3}) and $U$-string
(\ref{string3}) solutions.
All of the above transformations take each string into itself. We now
consider transformations that map one string into another. If we
compactify the $H$ action (\ref{H}) to {\it three} dimensions on $T^3$
the on-shell $SL(2,Z)_S$ will combine with the off-shell
$O(3,3;Z)$ target space duality to form an on-shell
$O(4,4;Z)$. Similar remarks apply to the $A$ and $B$ actions. It
follows that all three strings are mapped into one another by
$O(4,4;Z)$ transformations. That the {\it stringy cosmic string}
was related to the {\it elementary} string in this way was pointed out
in \cite{Sen7}; that the {\it dual string} was also related in this way
was pointed out in \cite{Rahmfeld2}.
\section{$N=2$ supergravity in $D=6$}
\label{sec:N2D6}
The preceding discussion has shown an interesting triality structure of the
$H$, $A$ and $B$ theories when compactified to four dimensions. However,
until now we have omitted the additional $D=6$ matter and/or gauge fields
present in all models. In this section we examine the full $D=6$, $N=2$
theories, and in the next section we incorporate the additional fields into
string/string/string triality.
We begin by focusing on the heterotic string compactified on a generic torus
to $D=6$ \cite{Narain,Narain2}.
The low-energy limit of this theory is described by a non-chiral
$N=2$ supergravity with one graviton multiplet and 20 Yang-Mills multiplets.
The bosonic action is given by
\begin{eqnarray}
I_H&=&\frac{1}{2\kappa^2}\int d^6x \sqrt{-G}e^{-\Phi}\Bigl[
R_G+G^{MN}\partial_M\Phi\partial_N\Phi
-\frac{1}{12}G^{MQ}G^{NR}G^{PS}H_{MNP}H_{QRS}\nonumber\\
&&\kern4em
+\frac{1}{8}G^{MN}{\rm Tr} (\partial_MML\partial_NML)
-\frac{1}{4}G^{MP}G^{NQ}F_{MN}{}^a(LML)_{ab}F_{PQ}{}^b\Bigr]\ ,
\la{eq:Hfull}
\end{eqnarray}
where $A_M{}^a$ are $24$ abelian gauge fields and
$H_{MNP}= 3(\partial_{[M}B_{NP]}+{\textstyle{1\over2}} A_{[M}{}^aL_{ab}F_{NP]}{}^b)$.
The 80 scalars parametrize an $O(4,20)/O(4)\times O(20)$ coset and
are combined into the symmetric $24\times24$ dimensional matrix $M$
satisfying $MLM=L$ where $L$ is the invariant metric on $O(4,20)$:
\begin{equation}
L=\pmatrix{0&I_4&0\cr I_4&0&0 \cr 0&0&-I_{16}}\ .
\la{L}
\end{equation}
The action is invariant under the $O(4,20;Z)$ target space duality
transformations $M\rightarrow\Omega M\Omega^T$,
$A_{\mu}{}^a\rightarrow\Omega^{a}{}_{b}A_{\mu}{}^b$,
$G_{\mu\nu}\rightarrow G_{\mu\nu}$, $B_{\mu\nu}\rightarrow B_{\mu\nu}$,
$\Phi\rightarrow\Phi$, where $\Omega$ is an $O(4,20;Z)$ matrix satisfying
$\Omega^TL\Omega=L$.
The full $I_H$ action is invariant under non-chiral six-dimensional
$N=2$ supersymmetry transformations. For convenience in writing down
fermionic equations, we use an underlying $D=10$ notation where the four
$D=6$ symplectic Majorana-Weyl spinors of the $N=2$ theory may be combined
into a ten-dimensional Majorana-Weyl spinor $\epsilon$. Since we will
need the supersymmetry transformations of the gravitino and dilatino when
deriving the Bogomol'nyi mass bound, we list them here:
\begin{eqnarray}
\delta\psi_M&=&\left[\nabla_M-{1\over8}H_{MNP}\Gamma^{NP}
+{1\over2\sqrt{2}} (VLF)_{R\,MN}^{\overline{a}}\Gamma^N\Gamma^{\overline{a}}
-{1\over4}(\partial_M^{\vphantom{-1}} V_R^{\vphantom{-1}}
V^{-1}_R)^{\overline{a}}{}_{\overline{b}}
\Gamma^{\overline{a}\overline{b}}\right]\epsilon\nonumber\\
\delta\lambda&=&-{1\over4\sqrt{2}}\left[\Gamma^M\partial_M\Phi
-{1\over6}H_{MNP}\Gamma^{MNP}+{1\over2\sqrt{2}}(VLF)_{R\,MN}^{\overline{a}}
\Gamma^{MN} \Gamma^{\overline{a}}\right]\epsilon\
\label{eq:hetsusy}
\end{eqnarray}
where the Dirac matrices may be given a ten-dimensional interpretation,
$\Gamma^{(10)}=\{\Gamma^A,\Gamma^{\overline{a}}\}$, with six-dimensional
Dirac matrices $\Gamma_M=E_M^A \Gamma^A$ \cite{Romansf4}.
Turning to the Type $IIA$ string compactified on $K3$, we find an identical
massless spectrum, corresponding to one $N=2$ supergravity multiplet coupled
to 20 $N=2$ Yang-Mills multiplets \cite{Nilsson2}. This time the action
is given by
\begin{eqnarray}
I_A&=&\frac{1}{2\kappa^2}\int d^6x \sqrt {-\tilde G}e^{-{\tilde \Phi}}\Bigl[
R_{\tilde G}+\tilde G^{MN}\partial_M{\tilde \Phi}\partial_N{\tilde \Phi}
-\frac{1}{12}\tilde G^{MQ}\tilde G^{NR}\tilde G^{PS}
\tilde H_{MNP}\tilde H_{QRS}\nonumber\\
&&\kern6em
+\frac{1}{8}{\tilde G}^{MN}{\rm Tr} (\partial_M{\tilde M}L\partial_N{\tilde M}L)
-{1\over4}e^{\tilde\Phi}\tilde G^{MP}\tilde G^{NQ}\tilde F_{MN}^a(L\tilde
ML)_{ab}\tilde F_{PQ}{}^b\Bigr]\nonumber\\
&&-{1\over2\kappa^2}\int d^6x{1\over16}\epsilon^{MNPQRS}\tilde B_{MN}\tilde
F_{PQ}{}^aL_{ab}\tilde F_{RS}{}^b\ ,
\la{eq:Afull}
\end{eqnarray}
where now ${\tilde H}$ has no Chern-Simons corrections, ${\tilde
H}=d{\tilde B}$. The action (\ref{eq:Afull}) has the same $O(4,20;Z)$
symmetry as (\ref{eq:Hfull}) \cite{Nilsson1}.
In particular, the matrix $\tilde M$ of scalars
satisfies the constraint $\tilde M L\tilde M=L$.
Under heterotic/Type $IIA$ duality we have the following dictionary
\cite{Duffstrong,Witten} relating the two sets of fields:
\begin{eqnarray}
{\tilde \Phi}&=&-\Phi\nonumber\\
{\tilde G}_{MN}&=&e^{-\Phi}G_{MN}\nonumber\\
{\tilde H}&=&e^{-\Phi}*H \\
\tilde A_M&=&A_M\nonumber\\
\tilde M&=&M\ .
\end{eqnarray}
This gives, in particular, the Type $IIA$ gravitino and dilatino
supersymmetry transformations
\begin{eqnarray}
\delta\tilde\psi_M&=&\Biggl[\tilde\nabla_M
-{1\over8}\tilde H_{MNP}\Gamma^{\hat7}\tilde\Gamma^{NP}\nonumber\\
&&\qquad-{1\over8\sqrt{2}} e^{\tilde\Phi/2}
(\tilde V L\tilde F)_{R\,NP}^{\overline{a}}
(\tilde\Gamma_M\tilde\Gamma^{NP}-4\delta_M{}^N\tilde\Gamma^P)
\Gamma^{\overline{a}} -{1\over4}
(\partial_M^{\vphantom{-1}} \tilde V_R^{\vphantom{-1}}
\tilde V^{-1}_R)^{\overline{a}}{}_{\overline{b}}
\Gamma^{\overline{a}\overline{b}}\Biggr]\tilde\epsilon\nonumber\\
\delta\tilde\lambda&=&{1\over4\sqrt{2}}\left[\tilde\Gamma^M\partial_M\tilde\Phi
+{1\over6}\tilde H_{MNP}\Gamma^{\hat7}\tilde\Gamma^{MNP}
-{1\over2\sqrt{2}}e^{\tilde\Phi/2}
(\tilde V L\tilde F)_{R\,MN}^{\overline{a}}
\tilde\Gamma^{MN} \Gamma^{\overline{a}}\right]\tilde\epsilon\
\end{eqnarray}
where $\Gamma^{\hat7}$ is the six-dimensional chirality operator with
eigenvalues $\pm1$.
Actually, (\ref{eq:Afull}) is not quite the action obtained by
compactifying $IIA$ supergravity on $K3$ which really has only $23$
vectors and one $3$-form potential $A_{MNP}$ \cite{Duffliu}; we have taken the
liberty of dualizing the $3$-form. Note that before dualizing
the off-shell symmetry is only
$O(3,19;Z)$.
Finally we consider the compactification of the Type $IIB$ theory on $K3$
\cite{Townsendk3}.
Since this theory is chiral in ten dimensions, it yields the chiral $N=2$
theory in six dimensions with
1 supergravity and 21 tensor multiplets. While this theory has
no covariant action, the equations of motion for the (anti)-self-dual
three-forms may be determined from the well-known properties of $K3$.
Details of this procedure are presented in the appendix. The resulting
equations have an on-shell $O(5,21,Z)$ invariance with $5\times 21=105$
scalars parametrizing the coset $O(5,21)/O(5)\times O(21)$. There are
$21+5=26$ chiral 3-forms, which we denote collectively as
$\skew4\tilde{\tilde H}_3^{i\pm}$, satisfying the (anti)-self-duality condition
\begin{equation}
\skew4\tilde{\tilde H}_3^{i\pm} = \skew3\tilde{\tilde\eta}_{ij}
*\skew4\tilde{\tilde H}_3^{j\pm}\ ,
\label{eq:IIBSD}
\end{equation}
with
\begin{equation}
\skew3\tilde{\tilde\eta}=\pmatrix{-1\cr&1\cr&&-1\cr&&&1\cr&&&&\eta_{ij}}\ .
\end{equation}
We have written $\skew4\tilde{\tilde H}_3^{i\pm}$ in a given order such that the
first 4 fields correspond to the self-dual and anti-self-dual components of
$H^{(1)}$ and $H^{(2)}$ (the ten-dimensional NS-NS and R-R 3-forms,
respectively). The remaining 22 chiral 3-forms come from the
compactification of the ten-dimensional self-dual 5-form field strength on
$K3$.
These chiral 3-forms as a set satisfy 26 Bianchi identities/equations of
motion
\begin{equation}
d\skew5\tilde{\tilde {\cal H}}_3^I=0\ ,
\end{equation}
where the two sets of 3-forms are related by a vierbein $\skew2\tilde{\tilde V}^i_J$
\begin{equation}
\skew5\tilde{\tilde {\cal H}}_3=(\skew2\tilde{\tilde L}^{-1})(\skew2\tilde{\tilde V}^{-1}) \skew4\tilde{\tilde H}_3^\pm\qquad
\skew4\tilde{\tilde H}_3^\pm=\skew2\tilde{\tilde V}\skew2\tilde{\tilde L} \skew5\tilde{\tilde {\cal H}}_3\ .
\end{equation}
The $O(5,21)$ matrix $\skew2\tilde{\tilde L}$ is given by
\begin{equation}
\skew2\tilde{\tilde L}=\pmatrix{-\sigma^1\cr&\sigma^1\cr&&d_{IJ}}\ ,
\label{eq:tildetildeL}
\end{equation}
and $\skew2\tilde{\tilde V}$ satisfies
\begin{equation}
\skew2\tilde{\tilde V}^{-1}=[\skew3\tilde{\tilde\eta}\skew2\tilde{\tilde V}\skew2\tilde{\tilde L}]^T\ .
\end{equation}
The explicit form for $\skew2\tilde{\tilde V}$ is given in the appendix.
The equations of motion for the bosonic fields of model $B$ are given by
\cite{RomansSD}
\begin{eqnarray}
\skew4\tilde{\tilde R}_{MN}-{\textstyle{1\over2}} \skew4\tilde{\tilde G}_{MN}\skew4\tilde{\tilde R}&=&
{1\over4}\skew4\tilde{\tilde H}_{MPQ}^{i\pm}\skew4\tilde{\tilde H}^{i\pm}{}_N{}^{PQ}\nonumber\\
&&+{\rm Tr} [\partial_M^{\vphantom{1}}\skew2\tilde{\tilde V}_R^{\vphantom{1}}\skew2\tilde{\tilde V}_L^{-1}
\partial_N^{\vphantom{1}}\skew2\tilde{\tilde V}_L^{\vphantom{1}}\skew2\tilde{\tilde V}_R^{-1}]
-{1\over2}\skew4\tilde{\tilde G}_{MN} {\rm Tr} [\partial_P^{\vphantom{1}}
\skew2\tilde{\tilde V}_R^{\vphantom{1}}\skew2\tilde{\tilde V}_L^{-1}
\partial^P\skew2\tilde{\tilde V}_L^{\vphantom{1}}\skew2\tilde{\tilde V}_R^{-1}]\nonumber\\
&&\kern-9.4em
\nabla^M(\partial_M^{\vphantom{1}}\skew2\tilde{\tilde V}_R^{\vphantom{1}}\skew2\tilde{\tilde V}_L^{-1})^{ij}
-(\partial_M^{\vphantom{1}}\skew2\tilde{\tilde V}_R^{\vphantom{1}}\skew2\tilde{\tilde V}^{-1}\skew3\tilde{\tilde\eta}
\partial^M\skew2\tilde{\tilde V}\Vtt_L^{-1})^{ij}
= {1\over6} \skew4\tilde{\tilde H}_{MNP}^{i+}\skew4\tilde{\tilde H}^{j-\,MNP}\nonumber\\
\skew4\tilde{\tilde H}_3^{i\pm}&=&\skew3\tilde{\tilde\eta}_{ij}*\skew4\tilde{\tilde H}_3^{j\pm}\nonumber\\
d\skew4\tilde{\tilde H}_3^{i\pm}&=&(d\skew2\tilde{\tilde V}\Vtt^{-1})_{ij}\skew4\tilde{\tilde H}_3^{j\pm}\ .
\end{eqnarray}
We note that the Type $IIB$ dilaton is included implicitly as one of the
scalars in $\skew2\tilde{\tilde V}$. Thus the equations of motion
are written above in a canonical framework.
The supersymmetric variation of the canonical gravitino is
\begin{equation}
\delta\skew5\tilde{\tilde\psi}_M^a
=\left[\delta^a{}_b\nabla_M+{1\over4}\skew4\tilde{\tilde H}_{MNP}^{i+}\Gamma^{NP}(T^i)^a{}_b
\right]\epsilon^b\ ,
\end{equation}
where the spinors $\epsilon^a$ are right-handed symplectic Majorana-Weyl
with $a$ labeling the $4$ of $Sp(4)\simeq SO(5)$. The five self-dual
3-forms transform as a vector of $SO(5)$ and the matrices $T^i$ satisfy the
$SO(5)$ Clifford algebra $\{T^i,T^j\}=2\delta^{ij}$. The (anti)self-duality
conditions are essential for the closure of the supersymmetry algebra
\cite{RomansSD}.
In order to gain a better understanding of model $B$, we may consider a
few special limits. If we set the R-R moduli to zero, then the vierbein
(given in the appendix) decomposes as
\begin{equation}
\skew2\tilde{\tilde V}^i{}_J=
\left[\matrix{{1\over\sqrt{2}}e^{\skew0\tilde{\tilde\Phi}}\cr
&{1\over\sqrt{2}}e^{\skew0\tilde{\tilde\Phi}}\cr
&&{1\over\sqrt{2}}e^{\rho/2}\cr
&&&{1\over\sqrt{2}}e^{\rho/2}\cr
\vphantom{{1\over\sqrt{2}}}
&&&&\!I_{22}}\right]\times\left[\matrix{\vphantom{{1\over\sqrt{2}}}
-1&-e^{-\skew0\tilde{\tilde\Phi}}\cr
\vphantom{{1\over\sqrt{2}}}1&-e^{-\skew0\tilde{\tilde\Phi}}\cr
\vphantom{{1\over\sqrt{2}}}&&1&-e^{-\rho}-{\textstyle{1\over2}}(b)^2&b^J\cr
\vphantom{{1\over\sqrt{2}}}&&-1&-e^{-\rho}+{\textstyle{1\over2}}(b)^2&-b^J\cr
\vphantom{{1\over\sqrt{2}}}&&0&-O^i{}_Ib^I&\!\!O^i{}_Kd^{KJ}}\right
\end{equation}
where $(b)^2=b^Ib^Jd_{IJ}$. This shows explicitly the factorization into
the dilaton and the $O(4,20)$ moduli space of $K3$ with torsion. Due to
the $D=10$ symmetry between $H^{(1)}$ and $H^{(2)}$, we may choose
to eliminate a different set of moduli, giving instead
\begin{equation}
\skew2\tilde{\tilde V}^i{}_J=
\left[\matrix{{1\over\sqrt{2}}e^{\skew0\tilde{\tilde\Phi}}\cr
&{1\over\sqrt{2}}e^{\skew0\tilde{\tilde\Phi}}\cr
&&{1\over\sqrt{2}}e^{\rho/2}\cr
&&&{1\over\sqrt{2}}e^{\rho/2}\cr
\vphantom{{1\over\sqrt{2}}}
&&&&\!I_{22}}\right]\times\left[\matrix{\vphantom{{1\over\sqrt{2}}}
-1&-e^{-\skew0\tilde{\tilde\Phi}}-{\textstyle{1\over2}}(b')^2&&&-b'^J\cr
\vphantom{{1\over\sqrt{2}}}1&-e^{-\skew0\tilde{\tilde\Phi}}+{\textstyle{1\over2}}(b')^2&&&b'^J\cr
\vphantom{{1\over\sqrt{2}}}&&1&-e^{-\rho}\cr
\vphantom{{1\over\sqrt{2}}}&&-1&-e^{-\rho}\cr
\vphantom{{1\over\sqrt{2}}}0&O^i{}_Ib'^I&&&\!\!O^i{}_Kd^{KJ}}\right]\
\end{equation}
where now the $b'^I$ are R-R moduli arising from $H^{(2)}$. This gives
a different decomposition of $O(5,21)$ into $O(1,1)\times O(4,20)$ and
hints at a symmetry under exchange of $\skew0\tilde{\tilde\Phi} \leftrightarrow\rho$ where
$\rho$ is the $K3$ breathing mode. In fact,
this is nothing but the underlying ten-dimensional $SL(2,Z)_X$
symmetry of the Type $IIB$ supergravity. This may be made clear by
eliminating the torsion moduli, $b^I=b'^I=0$. In this case the matrix
$\skew4\tilde{\tilde M} =\skew2\tilde{\tilde V}^T\skew2\tilde{\tilde V}$ may be written
\begin{equation}
\skew4\tilde{\tilde M}=\Omega\left[\matrix{{\cal M}_X\otimes{\cal M}_Y\cr
&H^I{}_Kd^{KJ}}\right]\Omega\ ,
\end{equation}
where $\Omega$ swaps entries 2 and 4. The matrices ${\cal M}_X$ and ${\cal
M}_Y$ are $SL(2,Z)$ matrices defined according to (\ref{eq:sl2mat}) where
\begin{eqnarray}
X&=&-\ell+ie^{-(\skew0\tilde{\tilde\Phi}-\rho)/2}\nonumber\\
Y&=&d+ie^{-(\skew0\tilde{\tilde\Phi}+\rho)/2}
\end{eqnarray}
($d$ is the single modulus arising from the ten-dimensional 4-form
potential). This shows a decomposition of $O(5,21)$ into $O(2,2)\times
O(3,19)$ with the last factor identified with the moduli of $K3$
surfaces of constant volume. Since $\skew0\tilde{\tilde\Phi}-\rho=\Phi^{(10)}$
is just the ten-dimensional dilaton, $X$ is exactly the field on which the
original $SL(2,Z)_X$ acts.
This last example may be further motivated by considering a truncated
version of model $B$ without self-dual fields. The reduction of the
original ten-dimensional 3-forms gives
\begin{equation}
I_B^{H^{(i)}}={1\over4\kappa^2}\int\left[e^{-\skew0\tilde{\tilde\Phi}}
H_3^{(1)}*H_3^{(1)}
+e^{-\rho}H_3^{(2')}*H_3^{(2')}\right]\ .
\label{eq:h1h2main}
\end{equation}
The $H^{(i)}$ are related to their counterparts in $D=10$ and are
explicitly
defined in the appendix. The on-shell symmetry of this version is the
$O(2,2;Z)$ subgroup of $O(5,21;Z)$
acting on the first four components.
One subgroup of this $O(2,2;Z)$ is the discussed $SL(2,Z)_X$.
Another interesting one is the $O(1,1;Z)\simeq Z_2$ acting on the
first two components. This transformation takes $H^{(1)}$ into
$e^{-\skew0\tilde{\tilde\Phi}} *H^{(1)}$ and $\skew0\tilde{\tilde\Phi}$ into
$-\skew0\tilde{\tilde\Phi}$ and is therefore a strong/weak duality transformation
for the Type $IIB$ string. This transformation is precisely the one
transforming the $T$-string into the $U$-string.
\section{Reduction to $D=4$}
\label{sec:N4D4}
When models $H$, $A$ and $B$ are reduced to four dimensions, they all give
rise to $D=4$, $N=4$ supergravities coupled to 22 Yang-Mills multiplets.
{}From the heterotic point of view, it is straightforward to compactify the
six-dimensional theory, given by (\ref{eq:Hfull}), to four dimensions on a
two-torus. The resulting bosonic action may be written
\begin{equation}
I_H^4={1\over16\pi G}\int d^4x\sqrt{-g}e^{-\eta}\left[R+(\partial\eta)^2
-{1\over12}{\cal H}_{\mu\nu\lambda}{}^2+{1\over8}{\rm Tr} (\partial \overline{M}
\overline{L}\partial\overline{M}\overline{L})-{1\over4}{\cal F}_{\mu\nu}{}^T
(\overline{L}\overline{M}\overline{L}){\cal F}_{\mu\nu}\right],
\end{equation}
where the four-dimensional variables are given by the standard dimensional
reduction techniques. In particular, the 28 gauge fields ${\cal A}_\mu$
arise two from the metric, two from the antisymmetric tensor and 24 from the
gauge fields in six dimensions. We group them together according to
\begin{equation}
{\cal A}_\mu=[A_\mu^i\quad \overline{B}_{i\mu}\quad \overline A_\mu]^T\ ,
\end{equation}
where
\begin{eqnarray}
\overline{A}_\mu&=&A_\mu-A_\mu^iA_i\nonumber\\
\overline{B}_{i\mu}&=&B_{i\mu}-A_\mu^jB_{ij}+{\textstyle{1\over2}}\overline{A}_\mu^TLA_i\ .
\end{eqnarray}
Note that the six-dimensional gauge fields are denoted by $A_\mu$ whereas
the metric $U(1)$'s always carry an index $i=4,5$. The scalars parametrize
an $O(6,22)/O(6)\times O(22)$ coset with metric
\begin{equation}
\overline{L}=\pmatrix{&I_2\cr I_2\cr&&L}
\end{equation}
and may be written in a vierbein form
\begin{equation}
\overline{V}
=\left[\matrix{{1\over\sqrt{2}}E^{-1}\cr&{1\over\sqrt{2}}E^{-1}\cr&&I_{24}}
\right]\times\left[\matrix{I_2&G+B-C&-A^T\cr I_2&-G+B-C&-A^T\cr
0&VLA&V}\right]\ ,
\end{equation}
where $C={\textstyle{1\over2}} A^TLA$ and $G$ and $B$ refer to the $4,5$ components of the
respective fields. The 3-form ${\cal H}$ is dual to the axion as given by
$(\ref{eq:Haxion})$ and may be written ${\cal H}_{\mu\nu\lambda}=
3(\partial_{[\mu}\overline{B}_{\nu\lambda]}+{\textstyle{1\over2}}{\cal A}_{[\mu}
\overline{L}{\cal F}_{\nu\lambda]})$ where
\begin{equation}
\overline{B}_{\mu\nu}=B_{\mu\nu}-A_\mu^iA_\nu^jB_{ij}-A_{[\mu}^i
(\overline{B}_{i\nu]}-A_i^TL\overline{A}_{\nu]})\ .
\end{equation}
It is of course no surprise that this theory has an explicit $O(6,22;Z)$
symmetry as expected from a direct compactification from ten dimensions on
$T^6$. In fact, the above four dimensional action could have been written
directly without the extra step of compactifying to six dimensions.
However, for string/string/string triality, it is enlightning to see
explicitly the compactification from $D=6$ to $D=4$. In particular, in the
absence of scalars $A_i$ originating from the six-dimensional gauge fields,
we find the simple split
\begin{equation}
\overline{V}={1\over\sqrt{2}}E^{-1}\left[\matrix{I_2&G+B\cr I_2&-G+B}\right]
\oplus V\ ,
\end{equation}
indicating the limit
\begin{equation}
{O(6,22)\over O(6)\times O(22)}\to
\left.{O(2,2)\over O(2)\times O(2)}\right|_{TU}
\times{O(4,20)\over O(4)\times O(20)}\ .
\end{equation}
Reduction of the Type $IIA$ theory on $T^2$ yields instead the
four-dimensional action
\begin{eqnarray}
I_A^4&=&{1\over16\pi G}\int d^4x\sqrt{-\tilde g}e^{-\tilde\eta}\biggl[
\tilde R+(\partial\tilde\eta)^2
-{1\over12} \tilde {\cal H}_{\mu\nu\lambda}{}^2
+{1\over4}({\rm Tr} (\partial\tilde G^{-1}\partial\tilde G)
+{\rm Tr} (\partial\tilde B\tilde{G}^{-1}\partial\tilde B\tilde{G}^{-1}))\nonumber\\
&&\kern10.4em +{1\over8}{\rm Tr} (\partial \tilde{M}L \partial \tilde{M}L)
-{1\over4}(\tilde F_{\mu\nu}^i\tilde{G}_{ij}\tilde F_{\mu\nu}^j
+\tilde {\cal H}_{\mu i\nu}\tilde{G}^{ij}\tilde {\cal H}_{\mu j\nu})
\biggr]\nonumber\\
&&+{1\over16\pi G}\int d^4x\sqrt{-\tilde g}e^{-\tilde\sigma}\left[
-{1\over2}{\rm Tr} (\partial\tilde A^TL\tilde ML\partial\tilde A\tilde G^{-1})
-{1\over4}\tilde{\cal F}_{\mu\nu}^T (L\tilde{M}L)\tilde{\cal F}_{\mu\nu}
\right]\nonumber\\
&&+{1\over16\pi G}\int d^4x\biggl[
-{1\over8}\epsilon^{ij}\tilde B_{ij}
\tilde{\cal F}_{\mu\nu}^TL*\tilde{\cal F}_{\mu\nu}
-{1\over2}\epsilon^{ij}
(\tilde{\cal H}_{\mu i\nu}-\tilde B_{ik} \tilde F^k_{\mu\nu})
\tilde A_j^TL *(\tilde{\cal F}_{\mu\nu}-{\textstyle{1\over2}}\tilde A_l \tilde F^l_{\mu\nu})
\nonumber\\
&&\kern6.55em -{1\over12}\epsilon^{\mu\nu\lambda\sigma}\epsilon^{ij}
\tilde{\cal H}_{\mu\nu\lambda}\tilde A_i^TL\partial_\sigma\tilde A_j
\biggr]\ .
\label{eq:IAfour}
\end{eqnarray}
As written, only $\tilde F_{\mu\nu}^i\equiv2\partial_{[\mu}^{\vphantom{i}}
\tilde A_{\nu]}^i$
are true field strengths ($\tilde{A}^{i}_{\mu}$ are the gauge fields arising
from the compactification of the metric $\tilde{G}$ as in (\ref{metricred})).
The other 2-forms, $\tilde{\cal H}_{\mu i\nu}$ and $\tilde{\cal F}_{\mu\nu}$,
are the shifted six-dimensional fields:
\begin{eqnarray}
\tilde{\cal H}_{\mu i\nu}&=& \tilde{H}_{\mu i\nu}
+2\tilde{A}^j_{[\mu}\partial_{\nu]}^{\vphantom{j}}\tilde{B}_{ij}
=2\partial_{[\mu}\overline{\tilde{B}}_{i\nu]}
+\tilde B_{ij}\tilde F_{\mu\nu}^j\nonumber\\
\tilde{\cal F}_{\mu\nu}&=& \tilde{F}_{\mu\nu}
\>\,+2\tilde{A}^j_{[\mu}\partial_{\nu]}^{\vphantom{j}}\tilde{A}_{j}
\ =2\partial_{[\mu}\overline{\tilde A}_{\nu]}
+\tilde A_i\tilde F_{\mu\nu}^i\ ,
\end{eqnarray}
where the four-dimensional gauge fields are
\begin{eqnarray}
\overline{\tilde B}_{i\mu}&=&\tilde B_{i\mu}-\tilde A_\mu^j\tilde B_{ij}
\nonumber\\
\overline{\tilde A}_\mu&=&\tilde A_\mu-\tilde A_\mu^i\tilde A_i\ .
\end{eqnarray}
$\tilde{\cal H}_{\mu\nu\lambda}$ is the three-form field strength with the
standard Bianchi identity arising from the metric and antisymmetric tensor
gauge fields:
\begin{eqnarray}
\tilde{\cal H}_{\mu\nu\lambda}&=&
3(\partial_{[\mu}\overline{\tilde B}_{\nu\lambda]}
+\tilde A_{[\mu}^i\partial_\nu^{\vphantom{i}}
\overline{\tilde B}_{i\lambda]}^{\vphantom{i}}
+\overline{\tilde B}_{i[\mu^{\vphantom{i}}}\partial_\nu^{\vphantom{i}}
\tilde A_{\lambda]}^i)\nonumber\\
\overline{\tilde B}_{\mu\nu}&=&\tilde B_{\mu\nu}
-\tilde A_{[\mu}^i\overline{\tilde B}_{i\nu]}^{\vphantom{i}}
-\tilde A_\mu^i\tilde A_\nu^j\tilde B_{ij}\ .
\end{eqnarray}
The duality map relating model $H_{STU}$ to model $A_{TSU}$ is given by
\begin{eqnarray}
&\hbox{metric}\hfil&\tilde g_{\mu\nu}=e^{\sigma-\eta}g_{\mu\nu}\nonumber\\
&\hbox{$U$ field}\hfil&\tilde G_{ij}=e^{\sigma-\eta} G_{ij}\nonumber\\
&\hbox{$S$--$T$ interchange}\hfil&\tilde\eta=\sigma\qquad
\tilde a=-{\textstyle{1\over2}}\epsilon^{ij}B_{ij}\nonumber\\
&&\tilde\sigma=\eta\qquad \tilde B_{ij}=-\epsilon_{ij}a\nonumber\\
&\hbox{metric gauge fields}\quad&\tilde A_\mu^i=A_\mu^i\nonumber\\
&\hbox{$H$ gauge fields}\hfil&\tilde {\cal H}_{\mu i\nu}=e^{\sigma-\eta}
\epsilon_i{}^j*{\cal H}_{\mu j\nu}\nonumber\\
&\hbox{$D=6$ fields}\hfil&\overline{\tilde A}_\mu=\overline A_\mu
\qquad \tilde A_i=A_i\qquad \tilde M=M\ ,
\label{dualmap}
\end{eqnarray}
where $\eta(\tilde{\eta})$ and $\s(\tilde{\s})$ are the dilatons/T-moduli
of the relevant theories.
When reduced to four dimensions, model $B$ loses its chirality and now
admits a Lagrangian formulation. Each six-dimensional three-form of
definite chirality reduces to a single $U(1)$ field strength and one
scalar. Thus the 28 four-dimensional gauge fields come two from the
reduction of the metric and 26 from $\skew5\tilde{\tilde {\cal H}}_3$. Prior to
the imposition of the self-duality conditions, the latter field strengths
are given by
\begin{equation}
\skew4\tilde{\tilde F}^a_{i\, \mu\nu}=2\partial_{[\mu} \overline{\skew4\tilde{\tilde B}^a}_{i\nu]} \qquad
\overline{\skew4\tilde{\tilde B}^a}_{i\mu}=\skew4\tilde{\tilde B}^a_{i\mu} -\skew5\tilde{\tilde A}_\mu^j\skew4\tilde{\tilde B}^a_{ij}\ ,
\end{equation}
where $i=4,5$. This gives a double counting which is eliminated by the
six-dimensional self-duality conditions, (\ref{eq:IIBSD}). Thus
\begin{equation}
\skew5\tilde{\tilde {\cal F}}_{i\,\mu\nu}^\pm=\epsilon_i{}^j \eta*\skew5\tilde{\tilde {\cal F}}_{j\, \mu\nu}^\pm\ ,
\end{equation}
where
\begin{equation}
\skew5\tilde{\tilde {\cal F}}_{i\,\mu\nu}^\pm=\skew2\tilde{\tilde V}(\skew4\tilde{\tilde F}_{i\, \mu\nu}+\skew4\tilde{\tilde B}_{ij}\skew4\tilde{\tilde F}_{\mu\nu}^j)\ .
\end{equation}
Reduction of the six-dimensional 3-form field equations then give
\begin{equation}
\nabla_\mu\left[\skew2\tilde{\tilde L}\skew4\tilde{\tilde M}_{ab}
(\skew4\tilde{\tilde F}_i^{b\,\mu\nu}+\skew4\tilde{\tilde B}^b_{ij}\skew4\tilde{\tilde F}^{j\,\mu\nu})
-\epsilon_i{}^j\skew4\tilde{\tilde B}^a_{jk}*\skew4\tilde{\tilde F}^{k\,\mu\nu}\right]=0\ ,
\end{equation}
which is a set of $2\times26$ equations and should be viewed as a
combination of both Bianchi identities and equations of motion. The
remaining equations of motion may similarly be reduced. We may then construct
a Type $IIB$ action which yields these equations of motion,
although there is some ambiguity in whether to choose
$p$-forms or their duals. The canonical choice is obtained by mirror
transformation
of the Type $IIA$ action, yielding the $B_{TUS}$ model. The duality map
relating $H_{SUT}$ to $A_{UST}$ is obtained by repeating (\ref{dualmap})
for the mirror-transformed heterotic string, and the $A_{UST}$ dilaton is then
$\rho$. The heterotic-Type $IIB$ dictionaries are then obtained by
performing mirror transformations on the Type $IIA$ strings.
{}From the conjectured six-dimensional heterotic/Type $IIA$ duality and
the connection between $IIA$ and $IIB$ via mirror symmetry it follows
that we have indeed a triality between all three strings in $D=4$;
beyond the simplified discussion of section (\ref{sec:N2D4}). However,
since $U$ and $T$
are embedded in the full $O(6,22;Z)$ whereas $S$ is not, the elegant
exchange symmetries $S/T$ and $S/U$ are destroyed. Note that the $A_{TSU}$
action (\ref{eq:IAfour}) has only $SL(2,Z)_U$
off-shell (besides the obvious $O(4,20;Z)$)
even though, as explained in the Introduction, the string
has also an $SL(2,Z)_S$. Similarly
the Type $B_{UTS}$ action has only $SL(2,Z)_T$
off-shell even though the Type $IIB$ string has also an
$SL(2,Z)_S$.
Consequently, none of the three actions is $SL(2,Z)_S$ invariant,
in contrast to the truncated $H,A,B$ actions discussed in section
(3). Since $SL(2,Z)_S$ is still a perturbative Type $IIB$ symmetry,
however, four-dimensional string/string/string triality still implies
the $S$-duality of the heterotic string.
\section{Bogomol'nyi Spectrum}
\label{sec:N4Bog}
We may derive the Bogomol'nyi mass bound in this theory by following a
Nester procedure \cite{Nester,Dabholkar,Harveyliu}. Since masses are
defined with respect to a canonical metric, it is convenient to work in
canonical variables (which we denote by a caret). From a supergravity
point of view, this mass bound
originates from the $N$-extended supersymmetry algebra with central charges
\cite{Wittenolive,Osborn}.
Thus we start by noting that, up to equations of motion, the supercharge
(parametrized by $\epsilon$) is given by
\begin{equation}
Q_{\epsilon}=\int\overline{\epsilon}\gamma^{\mu\nu\lambda}
\nabla_\nu\hat\psi_\lambda d\Sigma_\mu=\int\overline{\epsilon}
\gamma^{\mu\nu\lambda}\hat\psi_\lambda d\Sigma_{\mu\nu}\ .
\end{equation}
Therefore the anticommutator of two supercharges is
\begin{equation}
\{Q_\epsilon,Q_{\epsilon'}\}=\delta_\epsilon Q_{\epsilon'}
=\int N^{\mu\nu} d\Sigma_{\mu\nu}\ ,
\end{equation}
where $N^{\mu\nu}=\overline{\epsilon'}\gamma^{\mu\nu\lambda}
\delta_\epsilon\hat\psi_\lambda$ is a generalized Nester's form.
Just as the canonical Einstein metric is Weyl scaled by the dilaton
relative to the $\sigma$-model metric, the canonical gravitino is
shifted by the dilatino:
\begin{equation}
\hat\psi_\mu=e^{\eta/4}(\psi_\mu+\sqrt{2}\gamma_\mu\lambda)\ .
\end{equation}
Since the reduction of the six-dimensional supersymmetry transformations,
(\ref{eq:hetsusy}), gives
\begin{eqnarray}
\delta\psi_\mu&=&\left[\nabla_\mu-{1\over8}{\cal H}_{\mu\nu\lambda}
\gamma^{\nu\lambda}+{1\over2\sqrt{2}}(\overline{V}_R\overline{L}
{\cal F})_{\mu\nu}^{\overline a}\gamma^\nu\Gamma^{\overline{a}}
+\cdots\right]\epsilon\nonumber\\
\delta\lambda&=&-{1\over4\sqrt{2}}\left[\gamma^\mu\partial_\mu\eta
-{1\over6}{\cal H}_{\mu\nu\lambda}\gamma^{\mu\nu\lambda}
+{1\over2\sqrt{2}}(\overline{V}_R\overline{L}
{\cal F})_{\mu\nu}^{\overline{a}}\gamma^{\mu\nu}\Gamma^{\overline{a}}
+\cdots\right]\epsilon\ ,
\end{eqnarray}
Nester's form may be expressed as
\begin{eqnarray}
N^{\mu\nu}&=&\overline{\epsilon'}\gamma^{\mu\nu\rho}
\delta_{\epsilon}\hat\psi_\rho\nonumber\\
&=&\overline{\epsilon'}\gamma^{\mu\nu\rho}
\Bigl[\nabla_\rho+{1\over24}e^{-\eta}{\cal H}_{\eta\lambda\sigma}
(\gamma_\rho\gamma^{\eta\lambda\sigma}-3\delta_\rho{}^\eta
\gamma^{\lambda\sigma})\nonumber\\
&&\qquad\qquad\quad-{1\over8\sqrt{2}}e^{-\eta/2}
(\overline{V}_R\overline{L} {\cal F})_{\lambda\sigma}^{\overline{a}}
(\gamma_\rho\gamma^{\lambda\sigma} -4\delta_\rho{}^\lambda\gamma^\sigma)
\Gamma^{\overline{a}}+\cdots\Bigr]\epsilon\nonumber\\
&=&N_0{}^{\mu\nu}+{1\over2\sqrt{2}}e^{-\eta/2}\overline{\epsilon'}
(\overline{V}_R\overline{L}({\cal F}-i\gamma^5*{\cal F})^{\mu\nu}
)^{\overline{a}}\Gamma^{\overline{a}}\epsilon+\cdots\ .
\end{eqnarray}
In the last line, $N_0{}^{\mu\nu}$ is Nester's original expression
\cite{Nester}, which gives the ADM mass when integrated over the boundary
at spatial infinity
\begin{equation}
\overline{\epsilon'}P_\mu\gamma^\mu\epsilon={1\over4\pi G}
\int_{S^2_\infty}*N_0\ .
\end{equation}
Defining the charges by the asymptotic behavior of the gauge fields
\begin{equation}
{\cal F}_{0r}\sim {Q\over r^2}\qquad *{\cal F}_{0r}\sim {P\over r^2}\ ,
\end{equation}
the surface integral of Nester's form gives
\begin{equation}
{1\over4\pi G}\int_{S^2\infty}*N=\overline{\epsilon'}\left[P_\mu\gamma^\mu
+{1\over2\sqrt{2}G}e^{-\eta_0/2}(\overline{V}_R\overline{L}
(Q-i\gamma^5P))^{\overline{a}}\Gamma^{\overline{a}}\right]\epsilon\ .
\end{equation}
Either application of the supersymmetry algebra or explicit calculation
then insures that this expression must be non-negative (provided the
equations of motion are satisfied). From a four-dimensional $N=4$ point
of view, the Bogomol'nyi bound may then be written%
\begin{equation}
M\ge |Z_1|,|Z_2|\ ,
\end{equation}
where%
\footnote{These central charges have been noted independently by Cveti\v c
and Youm in \cite{Cvetic}. Note, however, that our Nester procedure does
not yield the extra charge constraint found in \cite{Cvetic} on the basis
of black hole solutions.}
\begin{equation}
|Z_{1,2}|^2={1\over(4G)^2}e^{-\eta_0}\left[Q_R^2+P_R^2\pm
2\left(Q_R^2P_R^2-(Q_RP_R)^2\right)^{\textstyle{1\over2}}\right]\ .
\end{equation}
The six right-handed electric charges are given by
\begin{equation}
Q_R^{\overline{a}}=\sqrt{2}(\overline{V}_R\overline{L}Q)^{\overline{a}}\ ,
\end{equation}
(and similarly for $P_R$).
This generalizes the Bogomol'nyi bound of \cite{Harveyliu}, which
holds only when the two central charges are identical, $|Z_1|=|Z_2|$.
Note that by using (\ref{eq:vlvr}), the square of
the right handed charges may be expressed as the $O(6,22;Z)$ invariant
combination
\begin{equation}
Q_R{}^2=Q^T\overline{L}(\overline{M}+\overline{L})\overline{L}Q\ .
\end{equation}
This allows us to write the central charges as
\begin{equation}
|Z_{1,2}|^2=\frac{1}{16G^2}
\left[\g_{ia} {\cal M}_{Sij}(\overline{M}+\overline{L})_{ab}
\g_{jb}
\pm \sqrt{(\g_{ia}\e_{ij}\g_{jb})
(\g_{kc}\e_{kl}\g_{ld})(\overline{M}+\overline{L})_{ac}
(\overline{M}+\overline{L})_{bd}}
\right]\ ,
\label{centralcharge}
\end{equation}
where the electric and magnetic charges have been combined into a single
$SL(2,Z)\times O(6,22;Z)$ vector
\begin{equation}
\gamma_{ia}=\pmatrix{\a_S^a \cr \b_S^a}=\pmatrix{
e^{-\eta_0}\overline{M}^{-1} & -a_{(0)} \overline{L} \cr
0 & \overline{L}}^{ab} \pmatrix{Q\cr P}^b.
\end{equation}
The first feature to notice is that they are manifestly $SL(2,Z)_S$
invariant which is of relevance for $S$-duality invariance of heterotic
string theory. It is a well-known fact \cite{Sen7} that the spectrum of
states in the short $N=4$ multiplets is $SL(2,Z)_S$ invariant. In that case
$|Z_1|=|Z_2|$ and we recover from (\ref{centralcharge}) the Schwarz-Sen
formula
\begin{equation}
M^2=\frac{1}{16G^2}
\g_{ia} {\cal M}_{Sij}(\overline{M}+\overline{L})_{ab}
\g_{jb}\ .
\label{schwsen}
\end{equation}
However, a
discussion for the intermediate multiplets was missing so far. The
masses of the states in those multiplets are given by $m={\rm Max}
(|Z_1|,|Z_2|)$. Due to the familiar nonrenormalization theorems the
central charges do not receive any quantum corrections which also implies
that the masses are not renormalized. $S$-invariance of
(\ref{centralcharge}) now gives the expected result that the full
supersymmetric mass spectrum has that property.
For the truncated set of fields considered in section (\ref{sec:N2Bog}), we
return to the notation of right-handed charges $Q_R$ and $P_R$. If only
charges 1 and 2 are active, the central charges then reduce to
\begin{eqnarray}
(4G)^2|Z_1|^2&=&e^{-\eta_0}[(Q_R{}^1+P_R{}^2)^2+(Q_R{}^2-P_R{}^1)^2]
\nonumber\\
(4G)^2|Z_2|^2&=&e^{-\eta_0}[(Q_R{}^1-P_R{}^2)^2+(Q_R{}^2+P_R{}^1)^2]\ .
\label{eq:H2Bog}
\end{eqnarray}
This corresponds to the mass bound (\ref{us}) of section (\ref{sec:N2Bog}),
and agrees with the formula of \cite{Kallosh,Kalloshpeet}.
Now we are ready to repeat the analysis of section (\ref{sec:N2Bog}) for the
various black hole types. Again we choose vanishing background.
For dilaton couplings $a=\sqrt{3}$ and $a=1$
the square root term vanishes which implies $|Z_1|=|Z_2|$ and
(\ref{centralcharge})
reduces to the Schwarz-Sen mass formula. It was shown in \cite{Rahmfeld1}
that both black holes satisfy that Bogomol'nyi bound and therefore preserve
1/2 of the supersymmetries in $N=4$.
What happens to the other two black holes when embedded in the $N=4$ theory?
For the $a=1/\sqrt{3}$ black hole with charge vectors as given in
section (\ref{sec:N2Bog}) (the additional 24 electric and 24 magnetic charges
are zero) we find $|Z_1|=3/4$ and $|Z_2|=1/4$. With the
knowledge that the mass was given by $m=3/4$ we conclude that this
state preserves only one supersymmetry in $N=4$. This also holds
for dilaton coupling $a=0$. Here we find $m=|Z_1|=1$, $Z_2=0$, leading to the
same supersymmetry structure. Both black holes are in intermediate multiplets
of the $N=4$ supersymmetry algebra. All four values of $a$ yield special
cases of the general solutions recently found in \cite{Cvetic}.
\bigskip
It is also instructive to examine the Bogomol'nyi mass bound from the model
$A$ point of view. In this case we start with the supersymmetry variation
of the four-dimensional Type $IIA$ gravitino
\begin{eqnarray}
\delta\skew5\hat{\tilde\psi}_\mu&=&\Bigl[\nabla_\mu+{1\over24}e^{-\tilde\eta}
\tilde {\cal H}_{\eta\lambda\sigma}\Gamma^{\hat7}
(\gamma_\mu\gamma^{\eta\lambda\sigma}-3\delta_\mu{}^\eta
\gamma^{\lambda\sigma})\\
&&\ +{1\over16}(e^{-\tilde\eta/2}(\tilde G_{ij}\tilde F_{\lambda\sigma}^j
+\tilde{\cal H}_{\lambda i\sigma}\Gamma^{\hat7})\Gamma^i
+\sqrt{2}e^{-\tilde\sigma/2}(\tilde V_R\tilde L
\tilde {\cal F})_{\lambda\sigma}^{\overline{a}}\Gamma^{\overline{a}})
(\gamma_\mu\gamma^{\lambda\sigma}-4\delta_\mu{}^\lambda\gamma^\sigma)
+\cdots\Bigr]\epsilon\ .\nonumber
\end{eqnarray}
This gives for Nester's expression
\begin{eqnarray}
\tilde N^{\mu\nu}&=&\tilde N_0{}^{\mu\nu}
+\overline{\epsilon'}\Biggl[{1\over4}e^{-\tilde\eta/2}
\Bigl((\tilde G_{ij}\tilde F^{j\mu\nu}+\epsilon_{ij}
*\tilde{\cal H}^{\mu j\nu})-i\gamma^5
(\tilde G_{ij}*\tilde F^{j\mu\nu}-\epsilon_{ij}
\tilde{\cal H}^{\mu j\nu})\Bigr)\Gamma^i
\nonumber\\
&&\kern10em+{1\over2\sqrt{2}}e^{-\tilde\sigma/2}
(\tilde V_R\tilde L(\tilde{\cal F}
-i\gamma^5*\tilde{\cal F})^{\mu\nu})^{\overline{a}}\Gamma^{\overline{a}}
\Bigr]\epsilon\ .
\label{eq:ANes}
\end{eqnarray}
This shows that, as far as the six-dimensional gauge fields are concerned,
the Type $IIA$ mass bound is identical to that of the Heterotic string.
Indeed, since the $S$--$T$ interchange is only applicable to the $6\to4$
fields, only their contributions to the Bogomol'nyi bound are modified.
{}From (\ref{eq:ANes}) we see that the four charges coming from the
compactification on $T^2$ enter into the mass formula in the combinations
\begin{eqnarray}
\tilde Q^a&=&\tilde Q_G^a+\epsilon^a{}_b\tilde P_B^b\nonumber\\
\tilde P^a&=&\tilde P_G^a-\epsilon^a{}_b\tilde Q_B^b\ ,
\end{eqnarray}
where $\tilde Q_G$ and $\tilde Q_B$ are defined by the asymptotic behavior
\begin{eqnarray}
\tilde E_i{}^a\tilde F^i_{0r}&\sim&{\tilde Q_G^a\over r^2}\nonumber\\
\tilde E^i{}_a\tilde{\cal H}_{0ir}&\sim&{\tilde Q_B^a\over r^2}
\end{eqnarray}
($E$ is the 4,5 components of the vierbein) and similarly for $\tilde P_G$
and $\tilde P_B$. The two central charges are then given by
\begin{equation}
|\tilde Z_{1,2}|^2={1\over(4G)^2}\left[\tilde{\cal Q}^2+\tilde{\cal P}^2
\pm2\left(\tilde{\cal Q}^2\tilde{\cal P}^2-(\tilde{\cal Q}
\tilde{\cal P})^2\right)^{1\over2}\right]\ ,
\end{equation}
where we have grouped the 6 electric charges according to
\begin{equation}
\tilde{\cal Q}=[e^{-\tilde\eta/2}\tilde Q^a\quad
e^{-\tilde\sigma/2}\tilde Q_R^{\overline{a}}]^T\ .
\end{equation}
The right-handed charges $\tilde Q_R^{\overline{a}}$ are related to the
charges carried by the six-dimensional gauge fields
\begin{equation}
\tilde Q_R^{\overline{a}}
=\sqrt{2}(\tilde V_R\tilde L\tilde Q_F)^{\overline{a}}\ ,
\end{equation}
and correspond exactly to their heterotic counterparts
($\tilde Q_R^{\overline{a}}=Q_R^{\overline{a}}$ for
$\overline{a}=6,\ldots9$). Analogous definitions hold
for $\tilde{{\cal P}}$.
For vanishing $\tilde Q_R^{\overline{a}}$, the central charges become
\begin{eqnarray}
(4G)^2|\tilde Z_1|^2&=&e^{-\tilde\eta_0}[(\tilde Q_R{}^1+\tilde P_R{}^2)^2
+(\tilde Q_R{}^2-\tilde P_R{}^1)^2]\nonumber\\
(4G)^2|\tilde Z_2|^2&=&e^{-\tilde\eta_0}[(\tilde Q_L{}^1-\tilde P_L{}^2)^2
+(\tilde Q_L{}^2+\tilde P_L{}^1)^2]\ ,
\label{eq:A2Bog}
\end{eqnarray}
where the $6\to4$ charges are grouped into the combination
\begin{eqnarray}
\tilde Q_R{}^a&=&\tilde Q_G^a+\tilde Q_B^a\nonumber\\
\tilde Q_L{}^a&=&\tilde Q_G^a-\tilde Q_B^a\ .
\end{eqnarray}
For the Type $IIB$ string, we once again start with the four-dimensional
gravitino variation
\begin{equation}
\delta\skew6\hat{\skew5\tilde{\tilde{\psi}}}_\mu=\left[\nabla_\mu
-{1\over16}\skew4\tilde{\tilde G}_{ij}\skew4\tilde{\tilde F}_{\lambda\sigma}^j
(\gamma_\mu\gamma^{\lambda\sigma}-4\delta_\mu{}^\lambda\gamma^\sigma)\Gamma^i
-{1\over16}\skew5\tilde{\tilde {\cal F}}_{i\,\lambda\sigma}^{a+}
(\gamma_\mu\gamma^{\lambda\sigma}+4\delta_\mu{}^\lambda\gamma^\sigma)
\Gamma^iT^a\right]P_R\epsilon\ .
\end{equation}
Since the spinors are chiral in six dimensions, we have explicitly
inserted the projection $P_R={1\over2}(1+\Gamma^{\hat7})
={1\over2}(1+\gamma^5\Gamma^{\hat3})$ into the above. Taking into account
the self-duality of $\skew5\tilde{\tilde {\cal F}}^+$, we arrive at
\begin{equation}
\skew4\tilde{\tilde N}^{\mu\nu}=\skew4\tilde{\tilde N}_0^{\mu\nu}
+\overline{\epsilon'}\left[
{1\over4}\skew4\tilde{\tilde G}_{ij}(\skew4\tilde{\tilde F}^{j\,\mu\nu}-i\gamma^5*\skew4\tilde{\tilde F}^{j\,\mu\nu})\Gamma^i
-{1\over4}(\skew5\tilde{\tilde {\cal F}}_i^{a+\,\mu\nu}
-i\gamma^5*\skew5\tilde{\tilde {\cal F}}_i^{a+\,\mu\nu})\Gamma^i
T^a\right]P_R\epsilon+\cdots\ .
\end{equation}
In this picture it is natural to define the Kaluza-Klein electric and
magnetic charges
\begin{equation}
\skew4\tilde{\tilde F}^i_{0r}\sim {\skew4\tilde{\tilde Q}^i\over r^2}\qquad
*\skew4\tilde{\tilde F}^i_{0r}\sim {\skew4\tilde{\tilde P}^i\over r^2}\ .
\end{equation}
For the remaining gauge fields, we may define the $2\times26$ charges
\begin{equation}
\skew5\tilde{\tilde {\cal F}}^{a+}_{i\,0r}\sim {\overline{Q}^a_i\over r^2}\ .
\end{equation}
Self-duality then gives the relation between ``electric'' and ``magnetic''
charges, $\overline{Q}_i^a= \epsilon_i{}^j\overline{P}_j^a$. With these
definitions, the central charges in model $B$ have the form
\begin{eqnarray}
|\skew4\tilde{\tilde Z}_{1,2}|^2&=&{1\over(4G)^2}\Biggl[
(\skew4\tilde{\tilde Q}^i+\epsilon^i{}_j\skew4\tilde{\tilde P}^j)^2
+2(\overline{Q}^a\!\cdot\overline{Q}^a
+\overline{P}^a\!\cdot\overline{P}^a)\nonumber\\
&&\quad\pm2\left(4(\skew4\tilde{\tilde P}\cdot\overline{P}^a
+\skew4\tilde{\tilde Q}\cdot\overline{Q}^a)^2
+2(\overline{Q}^a\!\cdot\overline{P}^b\overline{Q}^a\!\cdot\overline{P}^b
-\overline{Q}^a\!\cdot\overline{P}^b \overline{Q}^b\!\cdot\overline{P}^a)
\right)^{1\over2}\Biggr]\ .
\end{eqnarray}
The contractions denoted by $\cdot$ are over $i=4,5$ and are done with the
metric $\skew4\tilde{\tilde G}$.
For the truncated models of section (\ref{sec:N2D4}), only one of the
six-dimensional fields is active. In this case, the two central charges
reduce to
\begin{equation}
|\skew4\tilde{\tilde Z}_{1,2}|^2={1\over(4G)^2}\sum_{i=4,5}
\left[\skew4\tilde{\tilde Q}_i +\epsilon_i{}^j\skew4\tilde{\tilde P}_j\pm
(\overline{Q}_i+\epsilon_i{}^j\overline{P}_j)\right]^2\ .
\label{eq:BbogZ}
\end{equation}
As previously, we denote left- and right-handed charges (with the vierbein
removed) in the combinations
\begin{eqnarray}
\skew4\tilde{\tilde Q}_{R,L}&=&\skew4\tilde{\tilde E}\skew4\tilde{\tilde Q}
\pm\skew4\tilde{\tilde E}^{-1}\overline{Q}\nonumber\\
\skew4\tilde{\tilde P}_{R,L}&=&\skew4\tilde{\tilde E}\skew4\tilde{\tilde P}
\pm\skew4\tilde{\tilde E}^{-1}\overline{P}\ ,
\end{eqnarray}
so that the central charges of (\ref{eq:BbogZ}) may be written
\begin{eqnarray}
(4G)^2|\skew4\tilde{\tilde Z}_1|&=&(\skew4\tilde{\tilde Q}_R{}^1+\skew4\tilde{\tilde P}_R{}^2)^2
+(\skew4\tilde{\tilde Q}_R{}^2-\skew4\tilde{\tilde P}_R{}^1)^2\nonumber\\
(4G)^2|\skew4\tilde{\tilde Z}_2|&=&(\skew4\tilde{\tilde Q}_L{}^1+\skew4\tilde{\tilde P}_L{}^2)^2
+(\skew4\tilde{\tilde Q}_L{}^2-\skew4\tilde{\tilde P}_L{}^1)^2\ .
\label{eq:B2Bog}
\end{eqnarray}
Compared to (\ref{eq:H2Bog}) the charges have no dilaton prefactor since they
have been defined canonically. This completes the identification of the
central charges in all three models.
The central charges of the truncated theories, as given by (\ref{eq:H2Bog}),
(\ref{eq:A2Bog}) and (\ref{eq:B2Bog}), are summarized in Table 2.
Naturally, in the heterotic ($S$) language we verify the result of
\cite{Harveyliu} that only the right-handed charges contribute to the
central charges. From the Type $II$ point of view we find a democracy
between right- and left-handers. Each handedness goes along with one
central charge.
Naturally, the same result is obtained by dualizing the central charges
of the heterotic string. This implies that the dual of the $N=4$
heterotic string must be a Type $II$ string.
\begin{table}
$$
\begin{array}{cc}
\rm string&\rm central\ charge\\
&\\
S\hbox{--string}&Z_1{}^2=(Q_R{}^1+P_R{}^2)^2+(Q_R{}^2-P_R{}^1)^2\\
&Z_2{}^2=(Q_R{}^1-P_R{}^2)^2+(Q_R{}^2+P_R{}^1)^2\\
T\hbox{--string}&{\tilde Z}_1{}^2=(\tilde Q_R{}^1+\tilde P_R{}^2)^2
+(\tilde Q_R{}^2-\tilde P_R{}^1)^2\\
&{\tilde Z}_2{}^2=(\tilde Q_L{}^1-\tilde P_L{}^2)^2
+(\tilde Q_L{}^2+\tilde P_L{}^1)^2\\
U\hbox{--string}&{\skew4\tilde{\tilde Z}}_1{}^2=(\skew4\tilde{\tilde Q}_R{}^1
+\skew4\tilde{\tilde P}_R{}^2)^2
+(\skew4\tilde{\tilde Q}_R{}^2-\skew4\tilde{\tilde P}_R{}^1)^2\\
&{\skew4\tilde{\tilde Z}}_2{}^2=(\skew4\tilde{\tilde Q}_L{}^1+\skew4\tilde{\tilde P}_L{}^2)^2
+(\skew4\tilde{\tilde Q}_L{}^2-\skew4\tilde{\tilde P}_L{}^1)^2
\end{array}
$$
\caption{Central charges for the three theories. We have removed a
prefactor of $4G$ as well as the asymptotic value of the dilaton field.}
\label{table3}
\end{table}
Although the physical states of all three strings must be identical as a
condition for string/string/string triality, the interpretation of the
spectrum in terms of elementary versus solitonic excitations is different in
the heterotic and Type $II$ theories (in $D=4$ the $IIA$ and $IIB$ {\it
elementary}
massive spectra
have identical interpretations). In order to examine the elementary string
excitations, we set all magnetic charges to zero in the mass bound.
For the truncated heterotic theory, Table 2 gives
\begin{equation}
|Z_1|^2=|Z_2|^2={1\over(4G)^2}e^{-\eta_0}[(Q_R{}^1)^2+(Q_R{}^2)^2]\ ,
\end{equation}
which indicates that all Bogomol'nyi saturated elementary states in the
heterotic theory fall into short multiplets. For the NS sector of the
heterotic string, the mass formula for string states,
$M^2=L_0=\overline{L}_0$, becomes
\begin{eqnarray}
M^2={1\over{16G^2}}e^{-\eta_0}[(Q_L)^2+(N_L-1)]
&=&{1\over16{G^2}}e^{-\eta_0}[(Q_R)^2+(N_R-{\textstyle{1\over2}})]\nonumber\\
&=&|Z_1|^2+{1\over{16G^2}}e^{-\eta_0}[(N_R-{\textstyle{1\over2}})]\ ,
\end{eqnarray}
giving the well-known result that the elementary heterotic states saturating
the Bogomol'nyi bound must satisfy $N_R={1\over2}$ \cite{Sen6,Rahmfeld1}.
On the other hand, from a Type $II$ point of view, the central charges are
given by
\begin{equation}
|\tilde Z_1|^2={1\over(4G)^2}e^{-\tilde\eta_0}
[(\tilde Q_R{}^1)^2+(\tilde Q_R{}^2)^2]\qquad
|\tilde Z_2|^2={1\over(4G)^2}e^{-\tilde\eta_0}
[(\tilde Q_L{}^1)^2+(\tilde Q_L{}^2)^2]\ .
\end{equation}
Thus the elementary Type $II$ string excitations saturating the Bogomol'nyi
bound may fall in either short or intermediate representations depending on
whether $(\tilde Q_L)^2=(\tilde Q_R)^2$ or not. The Type $II$ string
mass formula in the NS-NS sector is%
\footnote{Space-time bosons in the R-R sector satisfy a similar equation.
While no elementary string states carry R-R charge, states from the R-R
sector may be charged under the NS-NS gauge bosons.}
\begin{eqnarray}
M^2&=&{1\over(4G)^2}e^{-\tilde\eta_0}[(\tilde Q_L)^2+(\tilde N_L-{\textstyle{1\over2}})]
={1\over(4G)^2}e^{-\tilde\eta_0}[(\tilde Q_R)^2+(\tilde N_R-{\textstyle{1\over2}})]
\nonumber\\
&=&|\tilde Z_2|^2+{1\over(4G)^2}e^{-\tilde\eta_0}
[(\tilde N_L-{\textstyle{1\over2}})]=|\tilde Z_1|^2+{1\over(4G)^2}
e^{-\tilde\eta_0}[(\tilde N_R-{\textstyle{1\over2}})]\ .
\end{eqnarray}
This indicates that Bogomol'nyi states are in short multiplets for
$\tilde N_L=\tilde N_R={1\over2}$ and intermediate multiplets for $\tilde
N_L>\tilde N_R={1\over2}$ or $\tilde N_R>\tilde N_L={1\over2}$.
\section{String and fivebrane solitons}
\label{sec:N4Sol}
When the full set of fields are included, one may once again find the
three string soliton solutions of section (3) but now the zero-mode
structures will be more complicated. Ideally, in fact, one would like
them to correspond to the worldsheet field content of the heterotic,
Type $IIA$ and Type $IIB$ superstrings.
That the Type $IIA$ theory in $D=6$ admits a soliton with the correct
heterotic zero-modes was discussed in \cite{Senssd,Harvey}. Just as we
found the 4-parameter deformation in section (5)
by making $O(2,1)/O(2) \times
O(2,1)/O(2)$ transformations on the neutral solution so we may find the
extra 24 parameters by making $O(20,1)/O(20) \times O(4,1)/O(4)$
transformations. When combined with the translation modes and their
fermionic partners, one finds in this way for the physical degrees of
freedom a total of $8$ right moving bosons, $8$ right moving fermions
and $24$ left moving bosons appropriate to the fundamental heterotic
string \cite{Senssd}. In fact, the same result may be obtained
\cite{Harvey,Townsendseven,Duffliu} by starting with the physical
zero modes of the Type $IIA$ fivebrane soliton in $D=10$ \cite{Callan2},
namely the $d=6$ chiral supermultiplet
$(B^-{}_{\mu\nu},\lambda^I,\phi^{[IJ]})$, and wrapping the fivebrane
around $K3$ \cite{Minasian}.
Finding the Type $II$ strings as solitons of the heterotic string is
more problematical, however. Although the zero modes associated with
the $4$ NS charges may be obtained in the same way, this is not true of
the $24$ RR charges since the fundamental Type $II$ strings do not
carry these charges \cite{Senssd,Harvey}. The problem of identifying
these zero modes is akin to the missing monopole problem
\cite{Gauntlett} and requires a better understanding of the
role of $K3$ in counting the dimension of the moduli space.
Since the Type $IIA$/heterotic duality admits a $D=10$ fivebrane
interpretation, one might expect the same to be true of Type $IIB$ now
that it has been included in the picture via four dimensional
string/string/string triality. However, in this case the critical
solitonic string found in $D=4$ does not seem to be related to the
$D=6$ string obtained by wrapping the $D=10$ fivebrane around $K3$
since this latter string appears not to be critical \cite{Townsendseven}.
This is in need of further study.
\section{Conclusion}
\label{sec:Conclusion}
{}From one point of view, four-dimensional string/string/string triality
seems a trivial extension of what we already knew: $D=6$ string/string
duality accompanied by mirror symmetry. Yet, as we have seen, it has
far-reaching consequences. $D=6$ string/string duality satisfactory accounts
for strong/weak coupling duality of the Type $IIA$ string in terms of
$SL(2,Z)_T$,
the target space duality of the heterotic string, but leaves a gap in
accounting for the converse, because $SL(2,Z)_S$ takes R-R fields
of Type $IIA$ into their duals. Four-dimensional string/string/string duality
fills this gap: $SL(2,Z)_S$ is guaranteed by $D=6$ general covariance
of the Type $IIB$ string. Moreover, since the conjectured $SL(2,Z)_X$
of the Type $IIB$ string is just a subgroup of the $O(6,22;Z)_{TU}$
target space duality of the heterotic string, we see that this triality also
accounts for this symmetry and hence for {\it all} the conjectured
non-perturbative symmetries of string theory.
\bigskip
\bigskip
\noindent
{\Large {\bf Acknowledgements}}
It is a pleasure to thank Ashoke Sen for useful conversations.
\bigskip
\bigskip
\noindent
{\Large {\bf Note Added}}
After the completion of this work, we became aware of a paper by
Girardello, Porrati and Zaffaroni \cite{PorrHet}, which also
displays the $D=4$ heterotic/$IIA$ dictionary and also discusses
the absence of a perturbative $T$-duality in the Type $IIA$ theory and
hence a gap in deriving $S$-duality of the heterotic string from
$D=6$ string/string duality \cite{Duffstrong} alone. However,
this gap is filled by the $D=4$
string/string/string triality of the present paper: $SL(2,Z)_S$
is guaranteed by $D=6$ general covariance of the Type $IIB$ string.
\newpage
|
\section{Introduction} \label{sec:introduction}
Given an initial data set for the gravitational field $(\Sigma,
g_{ab}, K_{ab})$ \cite{ids} associated with a Cauchy surface in an
asymptotically flat spacetime, can we tell whether gravitational
collapse has proceeded to such a point that a black hole has formed in
that spacetime? In principle, the answer is yes: Given a complete
description of the matter fields in the spacetime (i.e., their initial
data and evolution equations), then using Einstein's equation, evolve
the initial data to reconstruct the entire spacetime and then see
whether the spacetime has a nonempty black-hole region and, if so,
whether and where it intersects $\Sigma$. In practice, however,
carrying out this construction is a highly nontrivial task, even in
the vacuum case and even when done numerically \cite{Anninos94}.
While there is currently no simple algorithm for determining from an
initial data set whether a black hole has formed, if a future trapped
region exists in $\Sigma$, it must lie within the black-hole region,
provided the spacetime satisfies the null-convergence condition
\cite{ncc,Wald84,HawkingEllis73}. Recall that a closed subset $C$ of
$\Sigma$ having the structure of a three-manifold with smooth (or at
least $C^2$) boundary, bounded away from spatial infinity, is said to
be a future trapped region if the convergence of the future-directed
null geodesics orthogonal to $\partial C$ and outward directed (in the
sense that the projection of null geodesic tangent vectors on $C$ into
$\Sigma$ point outward from $C$) is non-negative everywhere on
$\partial C$ \cite{Wald84}. Denoting the induced metric on $\partial
C$ by $h_{ab}$, the mean extrinsic curvature \cite{mean} of $\partial
C$ in $\Sigma$ by $H$, and the extrinsic curvature of $\Sigma$ in the
spacetime by $K_{ab}$, $C$ is a future trapped region if
\begin{equation} \label{ftr}
H \le K_{ab}h^{ab}
\end{equation}
on $\partial C$. Likewise, the total future trapped region of
$\Sigma$ (being the closure of the union of all future trapped regions
in $\Sigma$) along with its boundary ${\cal A}$ (the future apparent
horizon, on which $H = K_{ab}h^{ab}$), is contained within the
black-hole region (under the same conditions as before)
\cite{apparent}. So, in initial data sets with a non-empty total
future trapped region, gravitational collapse has proceeded
sufficiently far so that black holes have formed.
Are there any simple conditions that guarantee the existence of a
future trapped region in an initial data set? Thorne's ``hoop
conjecture'' offers a test of this type: If a body of mass $M$ is
sufficiently compact so that a hoop of circumference $4\pi M$ can
encircle the body no matter how it is rotated there about, then the
body must be contained within a horizon \cite{Thorne72,Flanagan91}.
While a precise version of this conjecture remains to be proven,
Schoen and Yau have proven that an initial data set containing a
region $\Omega$ with sufficient matter density must contain a future
or past apparent horizon \cite{SchoenYau83}. This interesting result
has the slight weaknesses that its requirement on the matter content
is so strict that arbitrarily small vacuum regions in $\Omega$ are not
allowed and that, as a time-symmetric theorem, we cannot conclude that
the apparent horizon must be a future horizon. (Of course, this last
criticism can be avoided by restricting oneself to initial data
sets that do not contain past trapped regions.)\ \ Further, a number
of sufficient conditions have been found for spherically symmetric
initial data sets \cite{ss}. More recently, using Jang's equation and
its properties as established by Schoen and Yau \cite{SchoenYau81},
Eardley has recently provided a remarkably simple proof of the
following theorem \cite{Eardley95,Eardley93}.
{\it Theorem (Eardley)}. Fix an asymptotically flat initial data set
for the gravitational field $(\Sigma, g_{ab}, K_{ab})$ \cite{ids}
satisfying the dominant-energy condition. If there exists a compact
region $\Omega \subset \Sigma$ such that $K_{ab}(g^{ab} - n^a n^b)$ is
no less than the surface-to-volume ratio of $\Omega$ for all unit
vectors $n^a$ everywhere on $\Omega$, then $\Sigma$ must contain an
apparent horizon.
Were the apparent horizon a {\em future} apparent horizon, then, as
noted, the conditions of this theorem would guarantee that
gravitational collapse has proceeded sufficiently far that a
black-hole region has formed. Unfortunately, as given, the theorem
alone does not allow one to draw this conclusion as it suffers from
the same problem of Schoen and Yau's theorem \cite{SchoenYau83} in
that the possibility that that all such horizons will be past apparent
horizons has not been eliminated. However, the time-asymmetry in the
hypotheses of Eardley's theorem ($K^a{}_a$ is strictly positive on
$\Omega$ indicating that the region is collapsing ``on average'') is a
strong indication that there should be a future apparent horizon
somewhere in $\Sigma$.
By changing our viewpoint, Eardley's theorem suggests an alternative
argument having the advantage of producing a strengthened version of
the above theorem under weaker hypotheses. In particular, we can now
show that $\Sigma$ must in fact contain a future apparent horizon.
Further, this argument has an entirely geometric character, which is
to be compared to Eardley's argument, which, through the use of Jang's
equation, has an analytic character.
To begin, notice that the induced metric $h^{ab}$ on a two-surface
${\cal S} \subset \Sigma$ can be written as $h^{ab} = g^{ab} - n^a
n^b$, where $n^a$ is either of the two unit-normal vectors to ${\cal
S}$. Therefore, the hypothesis of the above theorem guarantees that,
on ${\cal S}$, $K_{ab}h^{ab}$ is bounded from below by the
surface-to-volume ratio of $\Omega$, which we denote by
$\sigma(\Omega)$. Notice that this bound is independent of the
two-surface in $\Omega$. This suggests that if there exists a region
in $\Omega$ (having the structure of a three-manifold with boundary)
whose boundary's mean extrinsic curvature $H$ is bounded above by
$\sigma(\Omega)$, then Eq.~(\ref{ftr}) would hold on the boundary, and
hence the region would be a future trapped region.
Does such a region exist in $\Omega$? The appearance of the
surface-to-volume ratio $\sigma$ in Eardley's theorem suggests that we
study this quantity as function on the collection of regions $C$ in
$\Omega$. Consider a region $C \subset \Omega$ that is ``nearly
degenerate'' in the sense that it is either flat like a pancake of
thickness $r$, thin like a cigar of radius $r$, or small like a sphere
of radius $r$. Then, we expect (as in the flat space case) that
$\sigma(C) \approx \text{const}/r$, for $r$ sufficiently small,
showing that regions that are nearly degenerate in the sense that they
are small in one or more dimensions have very large surface-to-volume
ratios. This suggests that there is some sufficiently well-behaved
region in $\Omega$ having minimal surface-to-volume ratio. In fact,
we conjecture that there always exists a region $\hat{C} \subset
\Omega$, having the structure of a differentiable manifold with
boundary, that minimizes $\sigma$ over such regions $C \subset
\Omega$. (See conjecture~1 in Sec.~\ref{sec:conjecture}.)\ \
Remarkably, it then follows that $H \le \sigma(\Omega)$ on any open
subset of $\partial\hat{C}$ where the surface is $C^2$, i.e., on the
portion having a well-defined and continuous extrinsic curvature.
(The proof of this fact is given in Sec.~\ref{sec:lemma1}.)\ \
Putting this all together, we have
\begin{equation} \label{proof1}
H \le \sigma(\Omega) \le K_{ab}h^{ab}
\end{equation}
on the open subset of $\partial\hat{C}$ that is $C^2$, where the first
inequality is a consequence of the minimizing property of $\hat{C}$
and the second follows by hypothesis. Were $\partial\hat{C}$
everywhere $C^2$, then $\hat{C}$ would be a future trapped region. As
explained in Sec.~\ref{sec:conjecture}, this is not always the case.
However, it is expected that $\partial\hat{C}$ is sufficiently
well-behaved so that Eq.~(\ref{proof1}) holds over a sufficiently
large subset of $\partial\hat{C}$ that $\hat{C}$ is indeed a future
trapped region, in the sense that it must lie in the black-hole region
of the spacetime. In particular, we conjecture that $\partial\hat{C}$
is everywhere $C^{2-}$ (see below) and $C^2$ everywhere except on a
closed set ${\cal Z}$ of measure zero. (See Sec.~\ref{sec:conjecture}
for the statement and discussion of this conjecture.)\ \ Therefore,
although Eq.~(\ref{ftr}) may not hold everywhere on $\partial\hat{C}$,
it does hold on $\partial\hat{C} \setminus {\cal Z}$, which, with the
fact that the surface is $C^{2-}$, is sufficient to guarantee the the
region is trapped. (See theorem~5 in Sec.~\ref{sec:trapped_notc2}.)\
\ This proves the following strengthened version of Eardley's theorem.
{\it Theorem 1.} Fix an initial data set for the gravitational field
$(\Sigma,g_{ab},K_{ab})$ \cite{ids} and fix a subset $\Omega \subset
\Sigma$ that is a compact three-manifold with $C^2$ boundary. If
$K_{ab}h^{ab}$ is no less than the surface-to-volume ratio of $\Omega$
for all rank~2 orthogonal projection maps $h^a{}_b$ \cite{projop}
everywhere on $\Omega$, then there exists a future trapped region in
$\Omega$, provided conjecture~1 (stated in Sec.~\ref{sec:conjecture})
holds for $(\Omega,g_{ab})$.
Denoting the eigenvalues of $K^a{}_b$ by $(k_1,k_2,k_3)$, ordered so
that $k_1 \le k_2 \le k_3$, it is worth noting that the minimum of
$K_{ab}h^{ab}$ over all rank~2 orthogonal projection maps $h^a{}_b$
\cite{projop} is precisely $k_1 + k_2$. Therefore, the sole condition
of theorem~1 is that the sum of the two lesser principal (extrinsic)
curvatures be no less than the surface-to-volume ratio of $\Omega$,
everywhere on $\Omega$.
We can now assert (assuming conjecture~1) that if such a region
$\Omega$ exists in a Cauchy surface of an asymptotically flat
spacetime \cite{Wald84} satisfying the null-convergence condition
\cite{ncc}, the spacetime must contain a black hole with the future
trapped region therein.
Comparing the two theorems, we see that while neither locates the
future apparent horizon, theorem~1 does tell us some subset of
$\Omega$, namely $\hat{C}$, is contained within the future apparent
horizon. Further, we see that theorem~1 dispenses with the asymptotic
and energy conditions that were needed by Eardley because of their use
in Schoen and Yau's analysis of Jang's equation.
The remainder of this work is organized as follows. In
Sec.~\ref{sec:lemma1}, we prove the lemma providing the bound $H \le
\sigma(\Omega)$ on $\partial\hat{C}$. In Sec.~\ref{sec:general}, we
review Eardley's argument and then present a strengthened version of
theorem~1. In Sec.~\ref{sec:conjecture}, we state and discuss the two
conjectures in Riemannian geometry needed for this work. In
Sec.~\ref{sec:trapped}, we offer a new proof that future trapped
regions are trapped, which is then modified to establish the same
result for our weaker notion a future trapped region, and then we
discuss the possibility of further extending the notion of a future
trapped region. Lastly, in Sec.~\ref{sec:discussion}, we discuss the
strengths and weakness of our results.
Our conventions are those of Ref.~\cite{Wald84} with the notable
exception that our sign convention for the extrinsic curvature of our
initial data surfaces is such that positive $K$ is associated with
collapse in the sense that it measures the {\em convergence} of
future-directed geodesic normals to the surface. On the other hand,
$H$ measures the {\em divergence} of the outward geodesic normals to
the surface of a region within an initial data surface.
Recall that a map between manifolds is said to be $C^{k-}$ if the the
mapping is $C^{k-1}$ and its $(k-1)$-order derivatives of the
functions defining the mapping are locally Lipschitz
\cite{HawkingEllis73,Choquet-Bruhat89}. Thus, a $C^{2-}$ embedded
surface is $C^1$ and the derivative of the embedding map is locally
Lipschitz.
It proves very convenient to make the following definitions. Given a
manifold $N$ (possibly with boundary), define ${\cal C}^k(N)$
[${\cal C}^{k-}(N)$] to be the collection of compact subsets of $N$
having the structure of a manifold with $C^k$ ($C^{k-}$) boundary.
It is useful to keep in mind that
\begin{equation}
{\cal C}^0(N) \supset {\cal C}^{1-}(N) \supset {\cal C}^1(N) \supset
{\cal C}^{2-}(N) \supset {\cal C}^2(N) \supset \cdots.
\end{equation}
Elements of ${\cal C}^k(N)$ and ${\cal C}^{k-}(N)$ need not be
connected, i.e., they can have many connected components. Further, if
$\Omega \in {\cal C}^k(N)$, then $\Omega \in {\cal C}^k(\Omega)$, and,
similarly, if $\Omega \in {\cal C}^{k-}(N)$, then $\Omega \in {\cal
C}^{k-}(\Omega)$.
Lastly, for a map $\phi: A \rightarrow B$, $\phi[A]$ denotes the image
of $A$ in $B$, $A \setminus B$ denotes the set of elements in $A$ that
are not in $B$, and $\overline{A}$ denotes the closure of $A$.
\section{Proof that $H \le \sigma(\Omega)$ on $\partial\hat{C}$}
\label{sec:lemma1}
Denote the surface area, volume, and surface-to-volume ratio of a
region $C \in {\cal C}^1(\Sigma)$ by $A(C)$, $V(C)$, and $\sigma(C) =
A(C)/V(C)$, respectively. More explicitly,
\begin{mathletters} \label{def_AV}
\begin{eqnarray}
A(C) & = & \int_{\partial C} \epsilon_{ab}, \\
V(C) & = & \int_{C} \epsilon_{abc},
\end{eqnarray}
\end{mathletters}
where $\epsilon_{abc}$ is the volume element constructed from $g_{ab}$
and $\epsilon_{ab}$ is the volume element constructed from the metric
$h_{ab}$ induced on $\partial C$ by $g_{ab}$.
{\it Lemma 1.} Fix a pair $(\Omega,g_{ab})$, where $\Omega$ is a
compact three-dimensional manifold with $C^1$ boundary and $g_{ab}$ is
a smooth Riemannian metric on $\Omega$. If $\hat{C} \in {\cal
C}^1(\Omega)$ is such that $\sigma(\hat{C}) \le \sigma(C)$ for all $C
\in {\cal C}^1(\Omega)$ and $O$ is an open subset of $\partial\hat{C}$
where the surface is $C^2$, then $H \le \sigma(\hat{C}) \le
\sigma(\Omega)$ on $O$, where $H$ is the mean extrinsic curvature of
$\partial\hat{C}$ \cite{mean}. If, further, $O$ is in the interior of
$\Omega$, then $H = \sigma(\hat{C})$ on $O$.
{\it Proof.} The idea of the proof is simple: We calculate $\sigma$
as a function along certain well-behaved curves in ${\cal
C}^1(\Omega)$ containing $\hat{C}$, calculate its derivative at
$\hat{C}$, and then use the fact that $\hat{C}$ minimizes $\sigma$ in
${\cal C}^1(\Omega)$.
Although there are many curves in ${\cal C}^1(\Omega)$, by which we
mean one-parameter family of regions $C_\lambda \in {\cal
C}^1(\Omega)$, for simplicity, we shall restrict ourselves to families
arising from a smooth deformation of a region $C \in {\cal C}^1(\Omega)$
in the sense that $C_\lambda = \phi_\lambda[C]$ for some one-parameter
family of maps $\phi_\lambda: \Omega \rightarrow \Omega$ such that
$\phi_\lambda$ is a diffeomorphism between $\Omega$ and $\phi_\lambda[
\Omega]$, with $\phi_0$ being the identity map on $\Omega$. Our
requirement that $\Omega$ and $\phi_\lambda[\Omega]$ be diffeomorphic
is sufficient to guarantee that $\partial(\phi_\lambda[C]) =
\phi_\lambda[\partial C]$, which makes the following calculations
easier than they would be otherwise.
A particularly simple class of such deformations, which is sufficient
for our purposes, are those associated with the flows of fixed vector
fields on $\Omega$ \cite{Choquet-Bruhat89}. [That is, given a fixed
vector field $\xi^a$, for $p \in \Omega$, $\phi_\lambda(p)$ is the
point along the integral curve of $\xi^a$ containing $p$ a parameter
distance $\lambda$ from $p$.]\ \ In order that these deformations be
well defined on all of $\Omega$ for some positive $\lambda$, it is
necessary to restrict ourselves to vector fields that are inward
pointing everywhere on $\partial\Omega$ (where we consider vectors
tangent to $\partial\Omega$ as inward pointing, so $\xi^k n_k \le 0$
everywhere on $\partial\Omega$, where $n^k$ is the unit outward normal
to $\partial\Omega$). Otherwise, a point $p \in \partial\Omega$ where
$\xi^a$ is strictly outward pointing would be mapped ``out of''
$\Omega$, and hence the deformation constructed from it would not be
defined for any positive $\lambda$, no matter how small. A
deformation $\phi_\lambda$ constructed from an inward pointing vector
field is well defined for all $\lambda \ge 0$ and is a diffeomorphism
between $\Omega$ and $\phi_\lambda[\Omega]$.
Fix any inward pointing vector field $\xi^a$ whose support intersects
$\partial C$ within $O$ and construct its one-parameter family of
deformations $\phi_\lambda$. Evaluating $A$ and $V$ on $C_\lambda =
\phi_\lambda[C]$ using Eqs.~(\ref{def_AV}), differentiating with
respect to $\lambda$, and then evaluating at $\lambda = 0$, we find
that
\begin{mathletters} \label{perturb}
\begin{eqnarray}
A'(C) & = & \int_{O} H (\xi^k n_k) \epsilon_{ab}, \\
V'(C) & = & \int_{O} (\xi^k n_k) \epsilon_{ab},
\end{eqnarray}
\end{mathletters}
where $H$ is the mean extrinsic curvature of $\partial C$ and $n^k$ is
the outward unit normal to $\partial C$. Differentiating the equality
$\sigma(C_\lambda) = A(C_\lambda)/V(C_\lambda)$, evaluating at
$\lambda = 0$, and using Eqs.~(\ref{perturb}), we find that
\begin{equation} \label{perturb_sigma}
\sigma'(C) = {1 \over V(C)} \int_{O}
\biglb(H-\sigma(C)\bigrb) (\xi^k n_k) \epsilon_{ab}.
\end{equation}
Using this equation, we now establish our bound on $H$.
We begin with the case where $O$ is in the interior of $\Omega$. Fix
any point $p \in O$. To show that $H(p) = \sigma(\hat{C})$, suppose,
for contradiction, that $H(p) > \sigma(\hat{C})$. Then, using the
facts that $p$ is in the interior of $\Omega$ and $H$ is continuous at
$p$, it is not difficult to show that there exists an open
neighborhood $N$ of $p$ and a vector field $\xi^a$ such that: (1) the
support of $\xi^a$ is $\overline{N}$; (2) $\partial\hat{C}$ is $C^2$
on $\overline{N} \cap \partial\hat{C}$; (3) $(\overline{N} \cap
\partial\Omega) = \emptyset$; (4) $\biglb(H-\sigma(\hat{C})\bigrb) >
0$ on $N \cap \partial\hat{C}$; (5) $(\xi^k n_k) < 0$ on $N \cap
\partial\hat{C}$. Notice that $\xi^a$, being zero on
$\partial\Omega$, is inward pointing, so the one-parameter family of
deformations $\phi_\lambda$ constructed from $\xi^a$ is defined for
all $\lambda \ge 0$. Using Eq.~(\ref{perturb_sigma}), we see that
$\sigma'(\hat{C}) < 0$, which is impossible as otherwise, for
sufficiently small $\lambda$, the region $\phi_\lambda[\hat{C}]$ would
have a smaller surface-to-volume ratio than $\hat{C}$. Similarly, if
$H(p) < \sigma(\hat{C})$, there exists an open neighborhood $N$ of $p$
and a vector field $\xi^a$ satisfying the above except with the
inequalities in (4) and (5) both reversed. Using
Eq.~(\ref{perturb_sigma}), we again find that $\sigma'(\hat{C}) < 0$,
which is again a contradiction. Therefore, $H(p) = \sigma(\hat{C})$,
as claimed.
Otherwise, fix any point $p \in O$. To show that $H(p) \le
\sigma(\hat{C})$, suppose, for contradiction, that $H(p) >
\sigma(\hat{C})$. Then, there exists an open neighborhood $N$ of $p$
and a vector field $\xi^a$ satisfying the above with (3) replaced by:
(3') $\xi^a$ is inward pointing on $\overline{N} \cap \partial\Omega$.
Again the one-parameter family of deformations $\phi_\lambda$
constructed from $\xi^a$ is defined for all $\lambda \ge 0$. Using
Eq.~(\ref{perturb_sigma}), we see that $\sigma'(\hat{C}) < 0$, which
is contradicts the minimality of $\sigma$ at $\hat{C}$. Therefore,
$H(p) \le \sigma(\hat{C})$, as claimed.
Lastly, that $\sigma(\hat{C}) \le \sigma(\Omega)$ follows simply from
the facts that $\hat{C}$ minimizes $\sigma$ over ${\cal C}^1(\Omega)$
and $\Omega \in {\cal C}^1(\Omega)$. This completes the proof of
lemma~1.~$\Box$
\section{Strengthening theorem~1}
\label{sec:general}
We begin with two definitions. First define the scalar field $\kappa$
on $\Sigma$ by setting
\begin{equation} \label{def_kappa}
\kappa(p) = \min_{h^a{}_b} (K_{ab}h^{ab}),
\end{equation}
for each $p \in \Sigma$, where the minimum is over the set of all
rank~2 orthogonal projection maps $h^a{}_b$ at $p$ \cite{projop}.
That is, $\kappa(p)$ is the sum of the two lesser principal
(extrinsic) curvatures at $p$. Second, for any continuous function
$f$ on $\Sigma$, define the function $W_f$ on ${\cal C}^1(\Sigma)$ by
setting
\begin{equation} \label{def_W}
W_f(C) = \int_C f \epsilon_{abc},
\end{equation}
for each $C \in {\cal C}^1(\Sigma)$. Note that with $f=1$, $W_1(C) =
V(C)$.
The idea behind the proof of theorem~1 was to use the properties of
the region that minimizes the surface-to-volume ratio $\sigma$ on
${\cal C}^1(\Omega)$. Noting that $\sigma = A/V = A/W_1$, one way to
proceed in generalizing theorem~1 is to analyze the properties of the
region that minimizes $A/W_\kappa$ on ${\cal C}^1(\Omega)$. Assuming
the relevant generalized version of conjecture~1 holds, it can be
shown that there is a future trapped region in $\Omega$ provided that
$A(\Omega)/W_\kappa(\Omega) \le 1$ and $\kappa$ is non-negative (and
not everywhere zero) on $\Omega$. However, such an argument must fail
if $\kappa$ is negative somewhere on $\Omega$ since, by choosing
regions with large area in regions where $\kappa$ is negative, we can
find $C \in {\cal C}^1(\Sigma)$ for which the ratio $A(C)/W_\kappa(C)$
is negative and as large as we wish (i.e., $A/W_\kappa$ has no finite
lower bound in this case). The fact that $\kappa$ cannot be even
slightly negative on small subsets of $\Omega$ makes this route
unattractive, so we take an alternative path suggested by the argument
Eardley used in proving his theorem.
\subsection{Eardley's argument}
Fix an asymptotically flat initial data set for the gravitational
field $(\Sigma, g_{ab}, K_{ab})$ \cite{ids} with sources satisfying
the dominant energy condition. Schoen and Yau have shown that such an
initial data set does not contain an apparent horizon (either future
or past) if and only if there exists a scalar field $f$ satisfying
Jang's equation everywhere on $\Sigma$ \cite{SchoenYau81}. Eardley's
argument is that certain initial data are inheritly incompatible with
the existence of a global solution of Jang's equation, and therefore a
(future or past) apparent horizon must be present. This argument goes
as follows.
Defining
\begin{equation} \label{defh}
h^a = {D^a f \over \sqrt{1 + D^m f D_m f}},
\end{equation}
where $D_a$ is the derivative operator on $\Sigma$ associated with
$g_{ab}$, Jang's equation takes the simple form
\begin{equation} \label{Jang}
D_a h^a = K_{ab}(g^{ab} - h^a h^b).
\end{equation}
Noting that $h^m h_m < 1$ everywhere, define the scalar field
$\tilde\kappa$ on $\Sigma$ by setting
\begin{equation} \label{def_kappat}
\tilde\kappa(p) = \inf_{|x|<1}\left( K_{ab}(g^{ab}-x^ax^b) \right),
\end{equation}
at each point $p \in \Sigma$, where the infimum is over all vectors
$x^a$ at $p$ with $x^m x_m < 1$. (By continuity, the value of
$\tilde\kappa$ is unchanged if we modify its definition by taking the
minimum over all vectors $x^a$ at $p$ with $x^m x_m \le 1$.)\ \ Using
the fact that for $x^a \neq 0$
\begin{eqnarray}
K_{ab}(g^{ab}&-&x^ax^b) \nonumber \\
& = & (1-x^mx_m)K_{ab}g^{ab} \nonumber \\
&& + (x^mx_m)K_{ab} \left( g^{ab} - x^a x^b /(x^mx_m) \right),
\end{eqnarray}
it is not difficult that show that
\begin{equation}
\tilde\kappa = \min(K^a{}_a, \kappa) \le \kappa
\end{equation}
at each point. Define the function $\tilde S$ on ${\cal C}^1(\Sigma)$
by setting
\begin{equation}
\tilde S(C) = A(C) - W_{\tilde\kappa}(C),
\end{equation}
for each $C \in {\cal C}^1(\Sigma)$.
If a global solution of Jang's equation exists, it follows that
\begin{equation} \label{Jang2}
D_a h^a \ge \tilde\kappa,
\end{equation}
everywhere on $\Sigma$. Integrating Eq.~(\ref{Jang2}) over any region
$C \in {\cal C}^1(\Sigma)$ and using the fact that $h^k n_k < 1$
everywhere on $\partial C$, where $n^k$ is the outward unit normal to
$\partial C$, we find that
\begin{equation}
\tilde S (C) > 0.
\end{equation}
That is, $\tilde S$ is a strictly positive function on ${\cal
C}^1(\Sigma)$. Therefore, if there exists a region $\Omega \in {\cal
C}^1(\Sigma)$ with $\tilde S (\Omega) \le 0$, a global solution of
Jang's equation cannot exist, and, thus, by Schoen and~Yau's results,
a (future or past) apparent horizon must be present within $\Sigma$.
In this argument, we see that the function $\tilde S$ on ${\cal
C}^1(\Sigma)$ arises rather naturally. This suggests that we should
attempt to strengthen the above result by showing that when there
exists $\Omega \in {\cal C}^1(\Sigma)$ with $\tilde S (\Omega) \le 0$
a future trapped region must exist within $\Omega$ (without the need
for any asymptotic or stress-energy conditions). However, it turns
out that we can do a little better using the function $S$ on ${\cal
C}^1(\Sigma)$, defined by
\begin{equation} \label{def_S}
S(C) = A(C) - W_\kappa(C),
\end{equation}
for each $C \in {\cal C}^1(\Sigma)$, rather than $\tilde S$. Since
$\tilde\kappa \le \kappa$, it follows that $S(C) \le \tilde S(C)$,
and, hence, if $\tilde S(\Omega) \le 0$, then $S(\Omega) \le 0$. So a
future trapped region theorem using $\tilde S$ follows from such a
theorem for $S$. (See theorem~3, below).
It is worth noting that for many initial data sets, $S$ and $\tilde S$
will coincide. Using the facts that $K^a{}_a = k_1 + k_2 + k_3$ and
$\kappa = k_1 + k_2$, it follows that $\tilde\kappa < \kappa$ if and
only if $K_{ab}$ is negative definite ($K_{ab}x^a x^b < 0$ for all
nonzero $x^a$), and $\tilde\kappa = \kappa$ otherwise. Therefore, if
$K_{ab}$ is nowhere negative definite on $\Sigma$, i.e., nowhere is
the surface positively contracting in all directions, then $S(C) =
\tilde S(C)$ for all $C \in {\cal C}^1(\Sigma)$.
\subsection{New argument}
Our first notable property of $S$ is that any region $\hat{C} \in
{\cal C}^2(\Sigma)$ that is a stationary point of $S$ is a future
trapped region.
{\it Theorem 2.} If $\hat{C} \in {\cal C}^2(\Sigma)$ is a stationary
point of $S$ (in the sense that $S'(\hat{C}) = 0$ for all smooth
variations of $C$), then $\hat{C}$ is a future trapped region.
{\it Proof.} Fix any region $C \in {\cal C}^2(\Sigma)$ and any open
subset $O$ of $\partial C$. Then, for all smooth vector fields
$\xi^a$ whose support intersects $\partial C$ within $O$
\begin{equation} \label{perturb_S}
S'(C) = \int_{O} (H-\kappa) (\xi^k n_k) \epsilon_{ab}.
\end{equation}
Therefore, repeating the argument used in lemma~1 and using the fact
that $\hat{C}$ is in the interior of $\Sigma$ (as $\Sigma$ has no
boundary), we find that $H = \kappa$ on $\partial\hat{C}$. However,
as $\kappa \le K_{ab}h^{ab}$ on $\partial\hat{C}$, where $h^{ab}$ is
the metric induced on $\partial\hat{C}$, $H \le K_{ab}h^{ab}$ on
$\partial\hat{C}$. Therefore, $\hat{C}$ is a future trapped
region.~$\Box$
Note that if $\hat{C} \in {\cal C}^2(\Sigma)$ is a local minimum of
$S$ in the sense that there is an open set $N \subset \Sigma$ such
that $S(\hat{C}) \le S(C)$ for all $C \in {\cal C}^2(\Sigma)$ with $C
\subset N$, then $\hat{C}$ is a stationary point of $S$. Further, for
momentarily static initial data sets ($K_{ab} = 0$ on $\Sigma$),
$\kappa = 0$ on $\Sigma$, so $S$ is simply the surface area of
$C$. Therefore, in this case, the problem of finding stationary points
of $S$ is exactly the problem of finding surfaces whose area is
stationary (in the sense of theorem~2), e.g., minimal two-surfaces
\cite{Osserman86}.
Since finding stationary points of $S$ is a difficult task, it is
desirable to have an alternate condition that guarantees the existence
of a future trapped region. Mimicking the proof of theorem~1, we fix
a region $\Omega \in {\cal C}^2(\Sigma)$ and then analyze the
properties of a region that minimizes $S$ on ${\cal C}^1(\Omega)$. If
$S(\Omega) > 0$, $S$ may not have a minimum on ${\cal C}^1(\Omega)$.
To see this, note that as there exist regions with arbitrarily small
surface areas and volumes, $\inf_{{\cal C}^1(\Omega)}(S) \le 0$. Yet,
for an initial data set with $\kappa \le 0$ (as is the case for a
maximal hypersurface), there is no region $\hat{C} \in {\cal
C}(\Omega)$ that attains the infimum (being zero) as any such region
necessary has $S(\hat{C}) \ge A(\hat{C}) > 0$. However, for any
region $\Omega$ with $S(\Omega) \le 0$, we conjecture that $S$ does
have a minimum on ${\cal C}^1(\Omega)$. (In fact, we conjecture that
the minimizing region $\hat{C}$ is a member of ${\cal C}^{2-}(\Omega)$
and has further nice differentiable properties. See conjecture~2 in
Sec.~\ref{sec:conjecture}.)\ \ The idea behind this conjecture is that
if $\inf_{{\cal C}^1(\Omega)}(S) < 0$ (which is guaranteed to be the
case if $S(\Omega) < 0$), a sequence of regions $C_i$ with $S(C_i)$
approaching this infimum cannot become degenerate in the sense that
their volumes go to zero or their areas become infinite, while if
$\inf_{{\cal C}^1(\Omega)}(S) = 0$, then $S(\Omega) = 0$, so $\Omega$
itself is a minimizing region. Note that $\inf_{{\cal
C}^1(\Omega)}(S)$ must be finite as
\begin{equation}
\inf_{{\cal C}^1(\Omega)}(S) \ge -\max_\Omega(\kappa)V(\Omega);
\end{equation}
a lower bound that holds even if $\kappa$ is negative somewhere on
$\Omega$. This is to be compared to the difficulty in establishing a
similar result for the surface-to-volume ratio function $\sigma$ and
lack of any finite lower bound on $A/W_\kappa$ when $\kappa$ is
negative somewhere on $\Omega$. Using these ideas, the following
theorem shows that if $S$ is not strictly positive on ${\cal
C}^2(\Sigma)$, then $\Sigma$ must contain a future trapped region.
{\it Theorem 3.} If $S(\Omega) \le 0$ for some $\Omega \in {\cal
C}^2(\Sigma)$, then there exists a future trapped region in $\Omega$,
provided conjecture~2 (stated in Sec.~\ref{sec:conjecture}) holds for
$(\Omega,g_{ab})$.
{\it Proof.} By conjecture~2, there exists $\hat{C} \in {\cal
C}^{2-}(\Omega)$ that minimizes $S$ on ${\cal C}^1(\Omega)$ and
further $\partial\hat{C}$ is $C^2$ on $\partial\hat{C} \setminus {\cal
Z}$, where ${\cal Z}$ is a closed set of measure zero. Therefore, for
all one-parameter family of deformations constructed from an inward
pointing vector field on $\Omega$ whose support intersects
$\partial\hat{C}$ where the surface is $C^2$, we have $0 \le
S'(\hat{C})$. Using Eq.~(\ref{perturb_S}), with $C = \hat{C}$ and
repeating the argument used in lemma~1, we find that $H \le \kappa$ on
$\partial\hat{C} \setminus {\cal Z}$. However, as $\kappa \le
K_{ab}h^{ab}$ on all of $\partial\hat{C}$, where $h^{ab}$ is the
metric induced on $\partial\hat{C}$, $H \le K_{ab}h^{ab}$ on
$\partial\hat{C} \setminus {\cal Z}$. Therefore, $\hat{C}$ is a
future trapped region.~$\Box$
Note that if $\kappa \le 0$ on $\Sigma$ (as is the case for maximal
hypersurfaces), then there is no region $\Omega$ meeting the condition
of theorem~3 as $S(\Omega) \ge A(\Omega) > 0$. Further, the condition
of Eardley's theorem and theorem~1 that $\sigma(\Omega) \le
\min_{\Omega}(\kappa)$ implies that $A(\Omega) \le
\min_{\Omega}(\kappa) V(\Omega) \le W_\kappa(\Omega)$ and, therefore,
$S(\Omega) \le 0$, which is the sole condition of theorem~3.
Therefore, theorem~3 is stronger than theorem~1, which is stronger
than Eardley's theorem.
It is interesting to note that $S(C)$ can be expressed as a pure
surface integral by introducing any vector field $\zeta^a$ on $\Sigma$
(or merely on $\Omega$) having the property that $D_a \zeta^a =
\kappa$, where $D_a$ is the derivative operator associated with the
metric $g_{ab}$. With this, we have
\begin{equation}
S(C) = \int_{\partial C} (1-\zeta^k n_k) \epsilon_{ab}.
\end{equation}
For instance, a particularly simple choice of $\zeta^a$ is that given
by taking $\zeta^a = D^a \phi$ for a scalar field $\phi$. Then,
$\phi$ must be a solution of Poisson's equation $D_a D^a \phi =
\kappa$ and can be fixed uniquely by fixing boundary data for $\phi$
on $\partial \Omega$ (e.g., $\phi=0$ on $\partial\Omega$) or a
boundary condition on $\phi$ at infinity (though whether this can
always be accomplished is more subtle). We will not pursue this
formulation any further here as nothing new seems to gained from this
viewpoint.
In theorems~1 and~3, we have restricted ourselves regions $\Omega$
with $C^2$ boundary for the sake of simplicity, and we expect that
both theorems hold under weaker conditions. It would seem that the
weakest differentiability condition that should be imposed is that for
which it makes sense for a region to speak of a region being future
trapped.
\section{Two Geometrical Conjectures}
\label{sec:conjecture}
The relevance of theorems~1 and~3 rests heavily upon the following two
conjectures, which we believe to be true.
{\it Conjecture 1.} Fix a pair $(\Omega,g_{ab})$, where $\Omega$ is a
compact three-dimensional manifold with $C^2$ boundary and $g_{ab}$ is
a smooth Riemannian metric on $\Omega$. There exists $\hat{C} \in
{\cal C}^{2-}(\Omega)$ such that $\sigma(\hat{C}) \le \sigma(C)$ for
all $C \in {\cal C}^1(\Omega)$. [In other words, $\sigma$ has a
minimum on ${\cal C}^1(\Omega)$ and a minimizing region is a member of
${\cal C}^{2-}(\Omega)$.]\ \ Further, $\partial\hat{C}$ is $C^2$
everywhere except on the closed set of measure zero given by
$\partial {\cal W}$, where ${\cal W} = (\partial\Omega \cap
\partial\hat{C})$ (and $\partial {\cal W}$ is constructed viewing
${\cal W}$ as a subset of either $\partial\Omega$ or
$\partial\hat{C}$).
{\it Conjecture 2.} Fix a triple $(\Omega,g_{ab},\kappa)$, where
$\Omega$ is a compact three-dimensional manifold with $C^2$ boundary,
$g_{ab}$ is a smooth Riemannian metric on $\Omega$, and $\kappa$ is a
smooth scalar field on $\Omega$. If $S(\Omega) \le 0$, then there
exists $\hat{C} \in {\cal C}^{2-}(\Omega)$ such that $S(\hat{C}) \le
S(C)$ for all $C \in {\cal C}^1(\Omega)$. [In other words, $S$ has a
minimum on ${\cal C}^1(\Omega)$ and a minimizing region is a member of
${\cal C}^{2-}(\Omega)$.]\ \ Further, $\partial\hat{C}$ is $C^2$
everywhere except on the closed set of measure zero given by $\partial
{\cal W}$, where ${\cal W} = (\partial\Omega \cap \partial\hat{C})$
(and $\partial {\cal W}$ is constructed viewing ${\cal W}$ as a subset
of either $\partial\Omega$ or $\partial\hat{C}$).
Note that although $\partial {\cal W}$ is by its definition a closed
subset of $\partial\hat{C}$, its being a set of measure does not
appear to be guaranteed as there exist boundaries of positive measure.
In conjectures~1 and~2, we have asserted that the surface
$\partial\hat{C}$ is a $C^{2-}$ submanifold that is almost everywhere
$C^2$. It is too much to expect that $\partial\hat{C}$ will be
everywhere $C^2$ as we expect a discontinuity in its mean extrinsic
curvature $H$ where $\partial\hat{C}$ ``first intersects''
$\partial\Omega$, i.e., on $\partial {\cal W}$. To see this, suppose
conjectures~1 and~2 are true. Then, write $\partial\hat{C}$ as the
disjoint union of three sets as follows
\begin{equation}
\partial\hat{C} = (\partial\hat{C} \setminus {\cal W})
\cup ({\cal W} \setminus \partial {\cal W})
\cup (\partial {\cal W}).
\end{equation}
As $\partial\hat{C} \setminus {\cal W}$ is in the interior of
$\Omega$, $H = \sigma(\hat{C})$ and $H = \kappa$ in conjectures~1
and~2, respectively. However, as $\partial\hat{C}$ coincides with
$\partial\Omega$ on the open set ${\cal W} \setminus \partial {\cal
W}$, $H$ will equal the mean extrinsic curvature of $\partial\Omega$
on ${\cal W} \setminus \partial {\cal W}$. Therefore, in general, we
expect that $H$ will suffer a discontinuity on $\partial {\cal W}$.
So, as $H$ will not always be $C^2$, $\partial\hat{C}$ will not always
be $C^2$. However, note that this argument suggests that the lack in
continuity in the second-order partial derivatives defining the
surface arise from mere jumps and not divergences. It is this
property that suggests that the surface is $C^{2-}$.
While we shall not attempt to do so here, conjectures~1 and~2 can
probably be proven using the ideas and techniques of geometric measure
theory \cite{gmt}. Very roughly, we consider a subset ${\cal
V}(\Omega)$ of ${\cal C}^{1-}(\Omega)$ whose members are sufficiently
well-behaved that they have finite volume and surface area (using the
Hausdorff measure). One then argues that $S$ is a continuous function
(in some natural topology) on ${\cal V}(\Omega)$ and that the subset
of ${\cal V}(\Omega)$ defined by those $C \in {\cal V}(\Omega)$ such
that $S(C) \le S(\Omega)$ is compact. It then follows immediately
that there is a region $\hat{C} \in {\cal V}(\Omega)$ that achieves
the minimal value of $S$ on this set. The last step would be to
establish that $\hat{C}$ is actually a member of ${\cal
C}^{2-}(\Omega)$ and $C^2$ on $\partial\hat{C} \setminus {\cal W}$
(and that ${\cal W}$ is a set of measure zero). We leave the task of
showing that these steps can actually be completed open for
investigation.
\section{Future trapped regions are trapped}
\label{sec:trapped}
Although there exists theorems showing that future trapped regions
must lie within the black hole region of the spacetime, the arguments,
as given, require that their surfaces be everywhere $C^2$
\cite{Wald84,HawkingEllis73}. Here, we show that the same result
holds for regions with boundaries that are not quite this smooth, and
so deserve to be called future trapped regions. To make our method of
proof clear, we first cover the case where the surface of the region
is everywhere $C^2$. After this, we modify the proof to accommodate
our more general regions. We then discuss the possibility of further
generalizations.
While our method of proof is similar to the existing proofs for smooth
regions, there is a notable difference in the final derived
contradiction. The Hawking and~Ellis argument ends with the
contradiction that the area of $\partial C$ is no less than the area
of $\partial J^+(C) \cap {\cal J}^+$, which, being at infinity, is
infinite. The Wald argument ends with the contradiction that the
future expansion of the null generators of $\partial J^+(C)$ is
nonpositive on $\partial C$ and yet positive near ${\cal J}^+$. Here,
we end with the contradiction that there are null generators of
$\partial J^+(C)$ extending beyond ${\cal J}^+$ that possess a point
conjugate to $\partial C$ on ${\cal J}^+$.
Actually, it should be noted that the Wald argument contains a slight
error in that the local cross-sections of ${\cal J}^+$ constructed
need not have the requisite differentiability properties in order that
nearby cross-sections of $\partial J^+(C)$ have strictly positive
future expansion. A simple counterexample is provided by a smooth
closed region $C$ in a flat spatial hypersurface $\Sigma$ in Minkowski
spacetime with the property that all of $C$ lies to one side of a flat
plane ${\cal P}$ in $\Sigma$ except for a closed region $\partial C
\cap {\cal P}$ having a non-empty interior (as a subset of ${\cal
P}$). Then, it is not difficult to see that the null generators of
$\partial J^+(C)$ having past endpoint on $\partial C \cap {\cal P}$
intersect ${\cal J}^+$ and have zero expansion everywhere. Of course,
the Wald argument can easily be fixed by introducing an area type
argument, or by adopting the method of theorem~4, which can be viewed
as such a fix as it has much of its inspiration from the Wald
argument.
Our notion of asymptotic flatness is that given in Ref.~\cite{Wald84}.
We denote the manifolds of the ``physical'' and ``unphysical''
spacetime by $M$ and $M'$, respectively. We remind the reader that $M
= M' \setminus (\overline{J^+(i^0)} \cup \overline{J^-(i^0)})$, where
$i^0$ is the point representing spatial infinity. Therefore,
$\partial M = (i^0 \cup {\cal J}^+ \cup {\cal J}^-)$, where ${\cal
J}^{\pm} = (\partial J^{\pm}(i^0)) \setminus i^0$ are future and past
null infinity.
Furthermore, the theorems we prove are for strongly asymptotically
predictable spacetimes \cite{Wald84}, which are simply those
asymptotically flat spacetimes for which there exists an open globally
hyperbolic subset $V$ of $M'$ containing $\overline{J^-({\cal J}^+)
\cap M}$ (where the closure is as a subset of $M'$). Note that
$\partial M \subset V$. It can be shown that all globally hyperbolic
asymptotically flat spacetimes are strongly asymptotically
predictable. Further, the globally hyperbolic asymptotic region $V$
can be chosen so that it contains all of $M$ and an asymptotically
flat Cauchy surface $\Sigma$ for $M$ together with spatial infinity
$i^0$ is a Cauchy surface for $V$. Therefore, the requirement that a
subset $C$ of $\Sigma$ be closed and bounded away from infinity (so
there exists a neighborhood of $i^0$ disjoint from $C$) is equivalent
to the condition that $C$ be closed as a subset of $\Sigma' = (\Sigma
\cup i^0)$.
\subsection{Regions whose surfaces are $C^2$}
\label{sec:trapped_c2}
{\it Theorem 4.} Fix a smooth strongly asymptotically predictable
spacetime \cite{Wald84} satisfying the null-convergence condition
\cite{ncc}. Let $\Sigma'$ be a smooth asymptotically flat Cauchy
surface for $V$ and let $C \subset (\Sigma' \cap M)$ be a future
trapped region in the sense that $C$ is a closed subset of $\Sigma'$,
$\partial C$ is $C^2$, and the convergence of the outward
future-directed null normals to $\partial C$ is everywhere
non-negative. Then, $(C \cap J^-({\cal J}^+)) = \emptyset$. [That
is, $C \subset (\Sigma' \cap B)$, where $B$ is the black-hole region
of the spacetime.]
{\it Proof.} In the following, all of our constructions are carried
out solely within the asymptotic globally hyperbolic region $V$.
Therefore, statements regarding the openness or closedness of sets
refer to these properties in $V$ alone. Since $C$ does not contain
$i^0$ (as $C$ is a subset of $M$), $J^+(C)$ does not contain $i^0$.
Further, $J^+(C)$ is closed, since $C$ is a closed subset of
$\Sigma'$. (See exercise~8 from chapter~8 of Ref.~\cite{Wald84}.)\ \
Therefore, there is a neighborhood of $i^0$ disjoint from $J^+(C)$.
Suppose, for contradiction, that $(C \cap J^-({\cal J}^+)) \neq
\emptyset$. Then, $(J^+(C) \cap {\cal J}^+) \neq \emptyset$, and,
hence, $(J^+(C) \cap I^+(i^0)) \neq \emptyset$. It then follows that
$(\partial J^+(C) \cap I^+(i^0)) \neq \emptyset$. To see this, fix
any point $p \in (J^+(C) \cap I^+(i^0))$. Then, as there exists a
timelike curve $\gamma$ from $i^0$ to $p$ [which must lie entirely
within $I^+(i^0)$] and there exists an open neighborhood of $i^0$
disjoint from the closed set $J^+(C)$, the curve $\gamma$ must leave
$J^+(C)$ and therefore intersect $\partial J^+(C)$, showing that
$(\partial J^+(C) \cap I^+(i^0)) \neq \emptyset$.
Recall that if $p$ is any point on a null generator of $\partial
J^+(C)$ whose past endpoint on $\partial C$ has an open neighborhood
on which $\partial C$ is $C^2$, there must not be a point conjugate to
$\partial C$ between $\partial C$ and $p$
\cite{Wald84,HawkingEllis73}. Pick a point $p \in (\partial J^+(C)
\cap I^+(i^0))$ and a null generator $\nu$ of $\partial J^+(C)$
containing $p$. Then $\nu$ cannot possess a point conjugate to
$\partial C$ in $M$ (with respect to either the physical or unphysical
metric) nor on ${\cal J}^+$ (with respect to the unphysical metric).
However, in the physical portion of the spacetime $M$, it follows from
the null Raychaudhuri equation and the null-convergence condition that
the (physical) future convergence of the null generators of $\partial
J^+(C)$ is not only non-negative on $\partial C$, it is non-negative
everywhere to the future \cite{Wald84,HawkingEllis73}. Furthermore,
if such a generator has positive convergence $\rho_0 > 0$ at some
point, then it must possess a conjugate point within an affine
parameter time $2/\rho_0$ thereafter, provided the generator can be
extended this far. Therefore, as $\nu$ is future complete in the
physical metric in $M$ (as it intersects ${\cal J}^+$), the (physical)
convergence along $\nu$ must be zero in $M$. Therefore, in the
infinitesimal sense, the physical area of a bundle of outgoing
future-directed null rays orthogonal to $\partial C$ is constant along
$\nu$ (in $M$). In terms of the unphysical metric, this area is that
given by the physical area multiplied by the square of the conformal
factor. As this conformal factor is zero on ${\cal J}^+$, it follows
that $\nu$ possesses a point conjugate to $\partial C$ where it
intersects ${\cal J}^+$ (with respect to the unphysical metric), which
is a contradiction.~$\Box$
\subsection{Regions whose surfaces are not quite $C^2$}
\label{sec:trapped_notc2}
The problem with the proof of theorem~4 when $\partial C$ is not
everywhere $C^2$ is that it may happen that because of our choice of
$p$ in the last paragraph, $\nu$ may have its past endpoint at a place
on $\partial C$ where the surface is not $C^2$, thus making the final
conjugate point argument inapplicable. When $\partial C$ is
everywhere $C^{2-}$ and $C^2$ on $\partial C \setminus {\cal Z}$,
where ${\cal Z}$ is a closed set of measure zero, although we do not
have complete freedom in what choice to make for $p$, it turns out we
can always find one so that the past endpoint of its associated null
generator has a neighborhood within $\partial C$ on which the surface
is $C^2$, i.e., its past endpoint is somewhere on $\partial C
\setminus {\cal Z}$. The idea is that it is impossible for only the
generators of $\partial J^+(C)$ with past endpoint on ${\cal Z}$ to
make it beyond ${\cal J}^+$ as there are not ``enough of them'' to
make up a ``local piece'' of $\partial J^+(C)$, as ${\cal Z}$ is a set
of measure zero in $\partial C$.
We capture this idea using the notion of Hausdorff measure \cite{gmt}.
On a differentiable manifold $N$ with Riemannian metric, for any two
points $a$ and $b$ in $N$, define $d(a,b)$ to be the greatest lower
bound on the lengths of $C^1$ curves in $N$ connecting $a$ to $b$ [so
$(N,d)$ is a metric space]. For any subset $S \subset N$, set
$\text{diam}(S) = \sup_{a,b \in S}(d(a,b))$. Then, for any subset $A
\subset N$ and numbers $k$ and $\delta > 0$, set
\begin{equation}
{\cal H}^k_\delta(A) = \inf \sum_j \nu_k
\left( {\text{diam}(S_j) \over 2} \right)^k,
\end{equation}
where $\nu_k$ is the volume of a unit-ball in flat ${\Bbb R}^k$ when
$k$ is a non-negative integer (so $\nu_0 = 1$, $\nu_1 = 2$, $\nu_2 =
\pi$, $\nu_3 = 4\pi/3$, etc.) and an arbitrary positive constant
otherwise, and where the infimum is taken over over all countable
coverings $\{ S_j \}$ of $A$ (i.e., $A \subset \cup_j S_j$) with
$\text{diam}(S_j) \le \delta$. With this, the ${\cal H}^k$-measure of a
set $A$ is defined as
\begin{equation}
{\cal H}^k(A) = \lim_{\delta \rightarrow 0} {\cal H}^k_\delta(A).
\end{equation}
This limit is well defined (though possibly infinite) as ${\cal
H}^k_\delta(A)$ is non-decreasing in $\delta$. It is worth noting
that if ${\cal H}^k(A) < \infty$ then ${\cal H}^m(A) = 0$ for all $m >
k$. It can be shown that if $A$ is an $k$-dimensional $C^1$ embedded
submanifold of $N$ with $k \le \dim(N)$, then ${\cal H}^k(A)$
corresponds to the usual ``volume'' of this submanifold. For
instance, in the case $\dim(N)=3$, ${\cal H}^1(A)$ is the length of a
1-dimensional submanifold $A$, ${\cal H}^2(A)$ is the area of a
2-dimensional submanifold $A$, and ${\cal H}^3(A)$ is the volume of a
3-dimensional submanifold $A$.
With this, we say a subset $A$ of a differentiable manifold $N$ has
${\cal H}^k$-measure zero if ${\cal H}^k(A) = 0$. It can be shown
that this notion is independent of which Riemannian metric is chosen,
and, therefore, whether a subset of a (paracompact) manifold has
${\cal H}^k$-measure zero is dependent solely upon the set. In the
case where $k = \dim(N)$, ${\cal H}^k$-measure zero is identical to
the usual Lebesgue notion of measure zero on a differential manifold.
Furthermore, if $f$ is a locally Lipschitz map from the manifold $N$
to another differentiable manifold, it follows that if ${\cal H}^k(A)
= 0$, then ${\cal H}^k(f[A]) = 0$.
Using these concepts, we can now prove that our generalized future
trapped regions are indeed trapped.
{\it Theorem 5.} Fix a smooth strongly asymptotically predictable
spacetime \cite{Wald84} satisfying the null-convergence condition
\cite{ncc}. Let $\Sigma'$ be a smooth asymptotically flat Cauchy
surface for $V$ and let $C \subset (\Sigma' \cap M)$ be a future
trapped region in the sense that $C$ is a closed subset of $\Sigma'$,
$\partial C$ is everywhere $C^{2-}$ and, on $\partial C \setminus {\cal
Z}$, $\partial C$ is $C^2$ and the convergence of the outward
future-directed null normals to $\partial C$ is non-negative, where
${\cal Z}$ is a closed set of measure zero. Then, $(C \cap J^-({\cal
J}^+)) = \emptyset$. [That is, $C \subset (\Sigma' \cap B)$, where
$B$ is the black-hole region of the spacetime.]
{\it Proof.} Suppose, for contradiction, that $(C \cap J^-({\cal
J}^+)) \neq \emptyset$. Then, using the same argument as in
theorem~4, it again follows that $(\partial J^+(C) \cap I^+(i^0)) \neq
\emptyset$. We claim that there exists $p \in (\partial J^+(C) \cap
I^+(i^0))$ with an associated null generator $\nu$ having past
endpoint on $\partial C \setminus {\cal Z}$, an open subset of
$\partial C$ where the surface is $C^2$. We show this by arguing that
there are not enough generators with past endpoint on ${\cal Z}$ to
make up $\partial J^+(C)$ in $I^+(i^0)$ as follows.
First, the subset $\tilde{\cal Z}$ of $\partial J^+(C)$ consisting of
those points with null generators having past endpoint on ${\cal Z}$
has ${\cal H}^3$-measure zero. To see this, denote by ${\cal K}$ the
subset of $TV$ (the tangent bundle associated with $V$) consisting of
all pairs $(p,k^a)$ where $p \in \partial C$ and $k^a$ is an outward
future-directed null vector normal to $\partial C$ at $p$. Using the
fact that $\partial C$ is $C^{2-}$, it follows that there exists a
locally Lipschitz map from $\partial C \times {\Bbb R}$ onto ${\cal K}
\subset TV$. Next, since $\partial J^+(C) \setminus C$ is generated
by null geodesics with past endpoint on $\partial C$ and
future-directed outgoing tangent vector normal to $\partial C$, we see
that $\partial J^+(C) \setminus C$ is a subset of the projection of
$\exp({\cal K})$ onto $V$ (where $\exp$ is the smooth diffeomorphism
from $TV$ to $TV$ defined by the geodesic flow on $TV$). As both
$\exp$ and the projection map are smooth, it follows that $\partial
J^+(C) \setminus C$ is a subset of the image of a subset of $\partial
C \times {\Bbb R}$ under a locally Lipschitz map. Therefore, since
${\cal Z} \times {\Bbb R}$ has ${\cal H}^3$-measure zero as a subset
of $\partial C \times {\Bbb R}$ (which follows from the fact that
${\cal Z}$ has ${\cal H}^2$-measure zero as a subset of $\partial C$)
and since $\tilde{\cal Z}$ is a subset of the image of a subset of
${\cal Z} \times {\Bbb R}$ under a locally Lipschitz map, it follows
that $\tilde{\cal Z}$ has ${\cal H}^3$-measure zero in $V$. (Note
that it is in the establishment of this result that we use the fact
$\partial C$ is $C^{2-}$ and not merely $C^1$.)
Next, pick any point $q \in (\partial J^+(C) \cap I^+(i^0))$ and an
open neighborhood $O$ of $q$ with $O \subset I^+(i^0)$. Using the
fact that $\partial J^+(C)$ is an achronal $C^{1-}$ embedded
three-dimensional submanifold of $V$ (see proposition~6.3.1 of
Ref.~\cite{HawkingEllis73}), it follows that $\partial J^+(C) \cap O$
has positive ${\cal H}^3$-measure. (To see this, note that we can
choose $O$ so that it is diffeomorphic to an open subset of ${\Bbb
R}^4$ with $\partial J^+(C) \cap O$ corresponding to the graph of a
$C^{1-}$ function of three variables.)\ \ Therefore, as the subset of
$\partial J^+(C)$ consisting of generators with past endpoint on
${\cal Z}$ has ${\cal H}^3$-measure zero, it follows that there must
exist a point $p \in \partial J^+(C) \cap O$ with an associated null
generator $\nu$ that has past endpoint on $\partial C \setminus {\cal
Z}$. (In fact, there are many such points.)
Arguing as we did in theorem~4 shows that $\nu$ contains a point
conjugate to $\partial C$ (with respect to the unphysical metric)
where $\nu$ intersects ${\cal J}^+$ (being between $\partial C$ and
$p$), which is a contradiction.~$\Box$
\subsection{Possible generalizations}
\label{sec:generalize}
In extending the notion of a future trapped region, we have restricted
ourselves to regions $C$ with $C^{2-}$ surfaces that are further $C^2$
everywhere except on a closed set of measure zero. We have done this
because this is both what we expect of the surfaces constructed
(conjectures~1 and~2) and these are regions for which we can carry
through all the relevant arguments (theorems~3 and~5). However, a
much greater extension seems possible. For instance, it is plausible
that the notion of a future trapped region can be extended to regions
with surfaces that are merely $C^{2-}$. Such a surface is twice
differentiable everywhere except on a set of measure zero ${\cal Z}$.
If the convergence of a family of future-directed outgoing null
geodesics orthogonal to a surface can be defined on $\partial C
\setminus {\cal Z}$ and the conjugate point argument used in theorem~5
can be applied to the generators with past endpoint on $\partial C
\setminus {\cal Z}$, the notion of a future trapped region with a
$C^{2-}$ surface would be a well-defined concept.
However, it would seem that the best notion of a region being future
trapped would not involve any differentiability conditions. For
example, consider the analogous problem of what we mean by a closed
region $C$ in flat space having a surface $S$ that is everywhere
locally convex. Here, we have a precise notion that imposes no
differentiability conditions on the surface: For each point $p \in S$
there is a neighborhood $N$ in $S$ such that $(1-\lambda) x + \lambda
y \in C$ for all $x,y \in N$ and $\lambda \in [0,1]$. (That is, the
convex hull of $N$ is a subset of $C$.)\ \ Likewise, we say the
surface of a region $C$ is locally concave if it is locally convex
when viewed as the surface of the closure of the complement of $C$.
Note that this flat space notion has a natural generalization to
curved spaces: We call the surface $S$ of a closed region $C$ locally
convex if for each point $p \in S$ there is a neighborhood $N$ in $S$
and a convex normal neighborhood $U$ containing $N$ such that for all
points $x, y \in N$ the geodesic from $x$ to $y$ (within $U$, being
unique) lies within $C$. In the $C^2$ case, the above implies the the
extrinsic curvature $H_{ab}$ of $S$ is positive semi-definite.
We want a geometric condition that, in the $C^2$ case, leads to the
bound $H = H^a{}_a \le K_{ab}h^{ab}$. Surely, such a notion would be
based on a demand that the areas of all local cross sections of
$\partial J^+(C)$ are non-increasing to the future (at least
sufficiently near $\partial C$). The problem is to capture this idea
in a well-defined sense. For instance, one needs for $\partial
J^+(C)$ to be sufficiently well-behaved so that the surface areas of
suitable cross-sections are well-defined. This is probably not such a
problem as $\partial J^+(C)$ is an imbedded $C^{1-}$ submanifold for
any set $C$. Then, to show that such regions are indeed trapped, an
area-type argument similar to that used by Hawking and~Ellis would
probably be the most natural method to use. However, how is one to
show that the areas of cross-sections are non-increasing to the future
when the null Raychaudhuri equation cannot be implemented? Clearly,
some subtlety is needed here.
Note that a naive condition such as $\partial C$ being everywhere
$C^{1-}$ and, on $\partial C \setminus {\cal Z}$, $\partial C$ is $C^2$
and the convergence of the outward future-directed null normals to
$\partial C$ is non-negative, where ${\cal Z}$ is a closed set of
measure zero, is insufficient. A simple counterexample is provided by
taking $C$ to be a solid cube in a flat spatial hypersurface in
Minkowski spacetime. Here, $\partial C$ is everywhere $C^{1-}$ and,
except along the edges and vertices (a closed set ${\cal Z}$ of
measure zero), the surface is $C^\infty$ and the convergence of the
outward future-directed null normals to $\partial C$ is zero.
However, $C$ is clearly ``visible'' from ${\cal J}^+$, i.e., it is not
trapped. In the proof of theorem~5, the problem with such surfaces is
that one does not have a one-to-one correspondence between the null
generators of $\partial J^+(C)$ and $\partial C$, and, as a result,
the portion of $\partial J^+(C)$ consisting of the generators having
past endpoint on ${\cal Z}$ has positive ${\cal H}^3$-measure. For
example, at a vertex, an entire ``octant's worth'' of null generators
of $\partial J^+(C)$ intersect $\partial C$ at a single point. In
this case, all null generators of $\partial J^+(C)$ that do make it
beyond ${\cal J}^+$ have past endpoints on ${\cal Z}$.
Lastly, one might expect that a differentiability condition that would
be sufficient to establish that a region $C$ is future trapped is that
$\partial C$ is everywhere $C^1$ and $C^2$ on an open dense subset $D$
of $\partial C$ (with the convergence of the outward future-directed
null normals being non-negative on $D$). In fact, this was the
approach first taken herein, but was abandoned because of a
difficulty. The idea is that if a null generator $\nu$ associated
with a point $p \in (J^+(C) \cap I^+(i^0))$ has its past endpoint on
$D$, the argument proceeds as in theorem~4, while if not, then it
would seem that we could find a point arbitrarily near $p$ in $(J^+(C)
\cap I^+(i^0))$ with an associated generator having past endpoint on
$D$. (After all, $D$ is dense in $\partial C$.)\ \ While this may be
true, proving it appears to be difficult. For instance, although one
might expect that there would exist a neighborhood of $\nu \cap
\partial C$ (within $\partial C$) such that all null generators with
past endpoint thereon remain on $\partial J^+(C)$ long enough to enter
$I^+(i^0)$, it turns out that this need not be the case if we just use
the fact that $\partial C$ is $C^1$. Whether this does hold when the
additional conditions on $\partial C$ are used is not clear.
\section{Discussion} \label{sec:discussion}
Theorems~1 and~3 provide us with simple tests for the existence of
future trapped regions within an initial data set, but how effective
are they?
First, the conditions of theorems~1 and~3 are quite strong in the
following sense. Recall that theorem~1 requires that
$\min_\Omega(\kappa) \ge \sigma(\Omega)$ (as does Eardley's theorem).
Using the fact that $\kappa$ is the sum of the two lesser principal
(extrinsic) curvatures $(k_1,k_2,k_3)$, it is not difficult to show
that $K^a{}_a = k_1+k_2+k_3 \ge {3 \over 2} \kappa \ge {3 \over 2}
\sigma(\Omega) > 0$ everywhere on $\Omega$, showing that this region
is everywhere contracting ``on average''. However, if $K^a{}_a$ is
non-negative, $\kappa$ need not be positive. This shows that the
region $\Omega$ is more than contracting ``on average''. Indeed, on a
maximal hypersurface ($K^a{}_a = 0$ everywhere on $\Sigma$), $\kappa
\le 0$ (with equality only where $K_{ab}=0$) everywhere on $\Sigma$.
In this respect, the condition of theorem~1 (and Eardley's theorem) is
quite strong. While theorem~3 merely requires that $S(\Omega) \le 0$,
so $\kappa$ need not be positive on all of $\Omega$, $\kappa$ still
must be positive over a sufficiently large subset of $\Omega$ in order
to meet this condition.
Second, while both theorems give sufficient conditions for the
existence of future trapped regions, neither condition is necessary.
This is easily seen by constructing a momentarily static initial data
set (so $K_{ab} = 0$, and hence $\kappa = 0$, on $\Sigma$) that
contains a minimal two-surface bounding a compact region $C$. This
region $C$ is future (and past) trapped and yet, as $S(C) = A(C)$ is
positive, the condition of neither theorem~1 nor~3 is met.
Third, neither theorem is very sensitive to the ``local''
existence of a future trapped region in the following sense. Suppose
we have a future trapped surface $S$ such that both families of
future-directed orthogonal null congruences have strictly positive
convergence on $S$. Construct a three-dimensional region $\Omega$ by
``thickening'' $S$ a small distance $r$ within an initial data surface
containing $S$. Then, for $r$ sufficiently small, $\Omega$ is a
future trapped region. However, for sufficiently small $r$,
$\sigma(\Omega)$ will be larger than $\inf_\Omega(\kappa)$ and
$S(\Omega)$ will be positive, and hence neither theorem enables us to
deduce that $\Omega$ itself is a future trapped region.
Fourth, the conditions of theorems~1 and~3 are quite robust in the
sense that if we have a region $\Omega$ that satisfies the condition
of either theorem with strict inequality and then deform it to create
a new region $\Omega'$ by ``pushing'' very thin fingers of the surface
of $\Omega$ into $\Omega$ (in arbitrarily complex ways), then,
provided our fingers are sufficiently thin, the surface area, volume,
and the integral of $\kappa$ for $\Omega'$ will be sufficiently near
those of $\Omega$ so that the conditions of both theorems will be met
for $\Omega'$. More generally, if we construct $\Omega'$ by excising
sufficiently thin regions from $\Omega$, both theorems guarantee the
existence of a future trapped region within $\Omega'$. This is
perhaps somewhat surprising at first given that $\Omega'$ can be
topologically quite complex. However, noting that the mean curvature
of the portions of the surface of $\Omega'$ created by excising ``very
thin fingers'' is very large and negative, we realize that $\hat{C}$
less the thin regions is nearly a future trapped region---all that is
needed is a bit of adjusting near the edges where the excised region
intersects $\hat{C}$.
Fifth, and last, theorems~1 and~3 do have a slight advantage in
numerical search for the existence of future trapped regions as
the calculation of $\sigma(\Omega)$ or $S(\Omega)$ requires only the
calculation of a surface area and a volume integral, which are not as
sensitive to numerical inaccuracies that would arise in calculating
the mean extrinsic curvature $H$ of a two-surface in $\Sigma$ to test
whether the surface is future outer trapped (i.e., testing whether the
condition given by Eq.~(\ref{ftr}) holds on the boundary).
So, while theorems~1 and~3 do offer tests for the existence of future
trapped regions, their inability to detect the existence of future
trapped regions in some instances, e.g., in initial data sets
associated with maximal hypersurfaces and ``thin'' future trapped
regions, leads us to wonder whether stronger tests of the type
considered here can be devised to give sufficient conditions for the
existence of future trapped regions.
\acknowledgements I thank Douglas Eardley, for his encouragement in
this work and for allowing me to visit the ITP at UCSB for its
discussion, and Robert Geroch, for answering numerous questions.
|
\section{\bf Introduction}
\setcounter{equation}{0}
For several last years an important progress has been achieved in
understanding the role of world-sheet superconformal symmetry and target
space symmetry of nonlinear $\sigma$-models in the context of string
theory and topological field theory \cite{W,Ler,Eg}. The BRST structure
of bosonic string ( $W_{n}$ string ) generates a topologically twisted
$N=2$ superconformal algebra \cite{GaSe} ($N=2$ super-$W_{n}$ algebra
\cite{BLNW,IK} ). In obtaining these results, a heavy use of the
hamiltonian reduction from WZNW models based on the superalgebra
$sl(n|n-1)$ has been made. Futhermore, any superstring theory possesses
$N=3$ twisted supersymmetry \cite{BLNW}. Recently, BRST structure has
been systematically constructed for superstrings with $N$ supersymmetries
by the hamiltonian reduction of affine extension of $osp(N+2|2)$ \cite{BLLS}.
$N=2$ analog for topological strings is the twisted $N=4$ $su(2)$
superconformal algebra (SCA) which has been obtained by the reduction of
affine extension of $sl(2|2)$ in \cite{LLS}.
As these and many other examples demonstrate, the hamiltonian reduction
is a powerful method of deducing new conformal \cite{DS,FRRTW,BTv} and
superconformal algebras and analysing the symmetry structure of the
conformal field theory and string theory models. Since a natural arena for
studying various superconformal symmetries and the related field theory
models is provided by superspace, it is tempting to have convenient
superspace generalizations of the hamiltonian reduction. $N=1$ superspace
version of this procedure in various aspects was discussed in Ref.
\cite{DRS}. On the other hand, a lot of interesting models (both in string
theory and topological field theory) reveal $N=2$ superconformal symmetries,
manifestly covariant formulations of which require $N=2$ superspace.
Motivated by this, in the present paper we generalize the hamiltonian
reduction procedure to $N=2$ superspace.
Let us recollect some well-known facts which are relevant to the problems
we address in the present paper.
Knizhnik \cite{K} and Bershadsky \cite{Ber} have proposed SCAs
with quadratic nonlinearity having as subalgebras $u(n)$ and
$so(n)$ affine algebras. It has been shown later \cite{GS} that the
nonlinear $so(3)$ and $so(4)$ Knizhnik-Bershadsky (KB) SCAs can be embedded
as subalgebras in usual linear $so(3)$ and $so(4)$ extended SCAs \cite{A}
after passing to some new basis for the currents of the latter (related to
the standard one by an invertible nonlinear transformation). By construction,
the usual $N=2$ and $N=4$ $su(2)$ SCAs \cite{A} are the same as $u(1)$ and
$u(2)$ KB SCAs, respectively.
Polyakov \cite{P} has found that there exist two types of classical
hamiltonian reductions for $sl(3)$: one yields $W_{3}$ algebra while the
other leads to $W_{3}^{(2)}$ which is a $u(1)$ "quasi" SCA in the sense
that dimension $3/2$ fields are bosonic ("wrong" statistics) and, besides,
it reveals a quadratic nonlinearity in the $u(1)$ current in its operator
product expansions (OPEs). Bershadsky \cite{B1} has further explained its
structure in detail. In Ref.\cite{Ro} new infinite
families of nonlinear extended conformal algebras, $u(n)$ and $sp(2 n)$
quasi SCAs, have been found. Independently it has been shown \cite{BTv,F}
that $u(n)$ quasi SCAs can be constructed by the hamiltonian reductions of
affine algebras $sl(n)^{(1)}$, based on non-principal embeddings of $sl(2)$
into $sl(n)$. A $N=2$ supersymmetric extension of
$W_3^{(2)}$ containing both $W_3^{(2)}$ and $N=2$ SCA as genuine subalgebras
have been constructed in \cite{{KrSo},{AKrSo}} by means of hamiltonian
reduction of the affine $sl(3|2)^{(1)}$ (at the level of component
currents). Recently, a formulation of this extended SCA in terms of
constrained $N=2$ superfields has been presented \cite{IKS}.
It was demonstrated in \cite{DTH,FL,Bo} that new SCAs
with quadratic nonlinearity, so-called $Z_{2}\times Z_{2}$ graded SCAs,
can be obtained by combining both fermionic and
bosonic spin-$3/2$ currents in the same $osp(m|2n)$ or $u(m|n)$
supermultiplet. The $u(n)$ KB SCAs and the algebra $W_{3}^{(2)}$ can be
identified with $Z_2\times Z_2$ graded SCAs associated with the
superalgebras $ u(n|0) $and $u(0|1)$, respectively
\footnote{There exist other conventions for these superalgebras, see,
e.g., Ref. \cite{FL}.}.
By applying the classical hamitonian reductuion to affine Lie superalgebra
$sl(n|2)^{(1)}$ and putting the constraint on the currents valued in its
bosonic $sl(2)$ part, in \cite{IM} the classical $u(n)$ KB SCAs has been
recovered in a new setting. In \cite{Ra}, this analysis was promoted to
$N=1$ superspace and a $N=1$ extension of $u(n)$ KB SCAs has been
constructed (at the classical level). However, an attempt to incorporate
$N=2$ supersymmetry has failed. As we will show, this happened just because
nonlinear constraints on $N=2$ affine supercurrents have not been involved
into the game.
As was already said, the aim of this paper is to develop the hamiltonian
reduction at the classical level directly in $N=2$ superspace. In short,
its main steps are: (i) construction of $N=2$ affine current algebra for
some superalgebra admitting a complex structure (we limit our consideration
here to the superalgebras $sl(n|n-1)$); (ii) imposing appropriate constraints
on the relevant superalgebra valued $N=2$ supercurrents; (iii) deducing $N=2$
extended superconformal algebras in $N=2$ superfield formalism.
We would like to specially emphasize that we are always dealing with $N=2$
superfield approach in our scheme. To our knowledge, this was not done
before. Another point to be mentioned is that our construction here is
purely algebraic and does not resort to any specific field theory
realization of $N=2$ affine current superalgebras, e.g. to their WZNW
realizations. This is the difference from, e.g., Ref. \cite{DRS} where
$N=1$ superspace version of the hamiltonian reduction was discussed in the
WZNW context. Also, we will be mainly interested in such extended $N=2$
SCAs which include as subalgebra the standard linear $N=2$ SCA, i.e.,
contain $N=2$ superconformal stress tensor among their defining
supercurrents.
The paper is organized as follows. In Section $2$ we construct
$N=2$ $sl(n|n-1)^{(1)}$ current algebra in terms of $N=2$ supercurrents
subjected to nonlinear constraints.
In Section $3$ we describe the general procedure of
the hamiltonian reduction in $N=2$ superspace and in Section 4 we exemplify
it by the simplest case of $N=2$
$sl(2|1)^{(1)}$ which gives rise to the standard $N=2$ SCA.
In Section $5$ we consider the case of $N=2$ $sl(3|2)^{(1)}$. We reproduce
the previously known $N=2$ $W_3$ and $N=2$ $W^{(2)}_3$ SCAs in
$N=2$ superfield formulation and find two
new $N=2$ extended SCAs. We explain how the
factorization of the dimension $1/2$ component currents in these
superalgebras
works. And finally in Section $6$ we end with a few closing remarks.
In Appendices,
we give notations for $sl(n|n-1)$ superalgebras, $u(m|n)$ SCA and different
realization of $sl(n|n-1)$.
\section{\bf $N=2$ Current Algebra for $sl(n|n-1)^{(1)}$ }
\setcounter{equation}{0}
In \cite{HS} Hull and Spence have constructed
$N=2$ current algebra for bosonic algebra $g$ in terms of
$N=2$ superfield currents satisfying {\it nonlinear} constraints.
The only essential restriction on $g$ is that it is
even-dimensional and admits a complex structure.
The quadratic terms appearing in the r.h.s. of
superoperator product expansions (SOPEs) between the supercurrents
are necessary for the consistency between these SOPEs and aforementioned
nonlinear constraints.
The nonlinearity of $N=2$ current algebra while it is written in terms
of $N=2$ supercurrents is the price for manifest $N=2$ supersymmetry.
When formulated via ordinary currents or $N=1$ supercurrents,
the algebra can be put in a linear form (in an appropriate
basis).
If $g$ is an ordinary bosonic algebra, all the $N=2$ affine supercurrents are
fermionic and we cannot put them to be constants. On the other hand,
this kind of constraints imposed on bosonic (super)currents
is of common use in the standard hamiltonian reduction scheme.
We are going to generalize the latter to $N=2$ superspace,
expecting such a generalization to allow us
to deduce extended $N=2$ SCAs (both previously known and new)
in a manifestly supersymmetric $N=2$ superfield fashion.
To be able to impose the aforementioned constraints on the affine
supercurrents, we need to have bosonic ones among them. A natural
way to achieve this is to deal with $N=2$ affine extensions of
{\it superalgebras}. So we are led to generalize the approach
of Ref. \cite{HS} to the superalgebras admitting a complex structure.
In this paper we
confine our consideration to the superalgebras $sl(n|n-1)$.
Let $g$ be a classical simple Lie superalgebra $ g=g_0 \oplus g_1$, where
$g_0$ is the bosonic subalgebra and $g_1$ is the fermionic subspace,
with the generators $t_{A}$ satisfying graded commutation relations
$ [ t_{A}, t_{B} \} ={F_{A B}}^{C} t_{C} $.
Let us introduce new structure constants,
${f_{AB}}^{C}=
(-1)^{(d_{A}+1)d_{B}} {F_{AB}}^{C}, $ where
for $ t_{A} \in g_{\alpha},
\alpha \in 0,1 $ we used the grading
$d_{A}=\alpha+1$. Therefore, ${f_{AB}}^{C}$ are
antisymmetric in the indices A,B when A,B correspond to bosonic
generators and
symmetric otherwise.
It is convenient to choose a complex basis for $g$,
so that its generators are labelled by $a$ and $\bar{a}$,
$a=1,2,\ldots$, $\frac{1}{2} \dim \, g = {1\over 2} ((2n-1)^2-1)\;$.
In this basis the complex structure associated with the second
supersymmetry has eigenvalue $+i$ on the generators $t_{a}$ and
$-i$ on the conjugated ones $t_{\bar{a}} (= t_{a}^{\dagger} )$.
The Killing metric $g_{a \bar{b}} $ is given by $ Str(t_{a}t_{\bar{
b}})$, $g_{a \bar{b}} $ being symmetric for the indices related to
bosonic generators
and antisymmetric otherwise.
Any index can be raised and lowered with $g^{a \bar{b}}$
and $g_{a \bar{b}}$.
The affine superalgebra $\hat{g}=sl(n|n-1)^{(1)}$ we deal with
in this paper has the equal
number $2n(n-1)$ of fermionic and bosonic supercurrents. For example, in
the fermionic $g$ valued supercurrent in the fundamental representation
${\cal J} \equiv {\cal J}_A t_B g^{AB}\;,$
top-left $n \times n$ and bottom-right $(n-1)
\times (n-1)$ matrix elements are fermionic,
so that $d_{a}, d_{\bar{a}}=1$.
Then the bosonic
supercurrents are entries of the top-right $n \times (n-1)$ and
bottom-left $(n-1)
\times n$ blocks in the supercurrent matrix, so for them
$d_{a}, d_{\bar{a}}=2$.
In the scheme of hamiltonian reduction which will be explained in the
next Section we impose non-zero constraints just on these supercurrents.
We refer the reader to Ref. \cite{HS} for details of how
the $N=2$ current algebra can be formulated in $N=2$ superspace.
The only new thing to be kept in mind in our case is that now
there are extra {\it bosonic} supercurrents besides the fermionic ones.
The presence of supercurrents with different statistics will play
an important role in our construction. This property will manifest
itself in the appearance of some extra $(-1)$ factors
in the r.h.s. of SOPEs defining the $N=2$ affine superalgebra.
With all these remarks taken into account, we summarize the
$N=2$ affine current algebra corresponding to
$sl(n|n-1)^{(1)}$ with the level $k$ as the following set of SOPEs
between $N=2$ superfield currents satisfying the appropriate nonlinear
constraints \footnote{ By Z we denote the coordinates of $1D$ $N=2$
superspace, $Z=(z, \theta, \bar{\theta})$. From now on we do not write
down explicitly the regular parts of SOPEs.
All the supercurrents (currents) appearing in
the r.h.s. of SOPEs (OPEs) are evaluated at the point
$Z_{2}$ ($z_2$).}:
\begin{eqnarray}
{{\cal J}}_{a} (Z_{1}) {{\cal J}}_{b} (Z_{2}) & = & -\frac{\bar{\theta}_{12}}{z_{12}} {f_{ab}}^{c} {{\cal J}}_{c}
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} (-1)^{(d_{a}+1)d_{\bar{c}}} (-1)^{(d_{b}+1)d_{\bar{c}}}
{f_{a \bar{c}}}^{
d} {f_{b}}^{\bar{c}e}
{{\cal J}}_{d} {{\cal J}}_{e}, \nonumber \\
{{\cal J}}_{\bar{a}} (Z_{1}) {{\cal J}}_{\bar{b}} (Z_{2}) & = &
\frac{\theta_{12}}{z_{12}} {f_{\bar{a} \bar{b}}}^{\bar{c}} {{\cal J}}_{\bar{c}}
+\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} (-1)^{(d_{\bar{a}}+1)d_{c}} (-1)^{(d_{\bar{b}}+1)d_{c}}
{f_{\bar{a}c}}^{\bar{d} }
{f_{\bar{b} }}^{c\bar{e}} {{\cal J}}_{\bar{d}} {{\cal J}}_{\bar{e}}, \nonumber \\
{{\cal J}}_{a} (Z_{1}) {{\cal J}}_{\bar{b}} (Z_{2}) & = &
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \frac{1}{2} k g_
{a \bar{b}} +\frac{1}{z_{12}} k g_{a \bar{b}}
+ \frac{\theta_{12}}{z_{12}} {f_{a \bar{b}}}^c {{\cal J}}_c - \frac{\bar{\theta}_{12}}{z_{12}}
{f_{a \bar{b}}}^{\bar{c}} {{\cal J}}_{\bar{c}} \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[ {f_{a \bar{b}}}^{c} \bar{{\cal D}} {{\cal J}}_{c}
- \frac{1}{k} (-1)^{(d_{a}+1)d_{\bar{c}}} (-1)^{(d_{\bar{b}}+1)d_{\bar{c}}}
{f_{a \bar{c}}}^{d} {f_{\bar{b}}}^{\bar{c} \bar{e}}
{{\cal J}}_{d} {{\cal J}}_{\bar{e}} \right],
\label{eq:qq}
\end{eqnarray}
where
\begin{eqnarray}
\theta_{12}=\theta_{1}-\theta_{2}, \; \bar{\theta}_{12}=\bar{\theta}_{1}-\bar{\theta}_{2}, \;
z_{12}=z_{1}-z_{2}+\frac{1}{2}(\theta_{1} \bar{\theta}_{2} + \bar{\theta}_{1} \theta_{2})
\end{eqnarray}
and the constraints on the supercurrents read
\begin{eqnarray}
{\cal D} {{\cal J}}_{a} - \frac{1}{2k} (-1)^{d_{a}} {f_{a}}^{bc} {{\cal J}}_{b} {{\cal J}}_{c} =
0, \quad
\bar{{\cal D}} {{\cal J}}_{\bar{a}} + \frac{1}{2k} (-1)^{d_{\bar{a}}}
{f_{\bar{a}}}^{\bar{b}
\bar{c}} {{\cal J}}_{\bar{b}} {{\cal J}}_{\bar{c}} = 0
\label{eq:cons}
\end{eqnarray}
(the summation is assumed over repeated indices).
Here, we work with complex fermionic covariant derivatives
\begin{eqnarray}
{\cal D}=\frac{\partial}{\partial \theta}-\frac{1}{2} \bar{\theta}
\partial,\;\;\;\;\;
\bar{{\cal D}} =\frac{\partial}{\partial \bar{\theta}}-\frac{1}{2}
\theta \partial \nonumber
\end{eqnarray}
satisfying the algebra
\begin{eqnarray}
\{ {\cal D}, \bar{{\cal D}} \}=-\partial\;(=-\partial_{z}),
\label{eq:DDB}
\end{eqnarray}
all other anticommutators are vanishing.
If we restrict the indices in (\ref{eq:qq}) and (\ref{eq:cons})
to the fermionic supercurrents we reproduce the $N=2$ $sl(n)^{(1)} \oplus
sl(n-1)^{(1)} \oplus u(1)^{(1)}$
affine current algebra \cite{HS}.
We have checked that the whole $N=2$ current superalgebra
(\ref{eq:qq}) with the nonlinear
constraints (\ref{eq:cons}) satisfies the standard $Z_2$ graded
Jacobi identities and that SOPEs of the l.h.s. of (\ref{eq:cons})
with any affine supercurrent vanish on the shell of constraints (the
presence of nonlinear terms in the r.h.s. of (\ref{eq:qq}) is crucial
for this).
When we consider this superalgebra at the quantum level (to all orders
in contractions between the supercurrents), then there appears an extra term,
$\frac{1}{2} (-1)^{d_{a}+1} {f_{a \bar{c}}}^{d} {f_{\bar{b}}}^{\bar{c}
\bar{d}}$ in $\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2}$ in the r.h.s. of SOPE ${\cal J}_{a}(Z_{1})
{\cal J}_{\bar{b}}(Z_{2})$.
This is due to the fact that there exist additional contractions
between the supercurrents at the quantum level. In the remainder of
this paper we will deal with the classical relations (\ref{eq:qq}) and
(\ref{eq:cons}).
Generalizing the well-known
Sugawara construction to $N=2$ superspace yields the following
formula for the improved $N=2$ stress
tensor in terms of the affine supercurrents ${\cal J}_{a}, {\cal J}_{\bar{a}}$,
\begin{eqnarray}
{{\cal T}}_{sug}=\frac{1}{k} g^{a \bar{b}} {\cal J}_{a} {\cal J}_{\bar{b}}
+\alpha_{i} \bar{{\cal D}} {\cal H}_{i}+\alpha_{\bar{j}} {\cal D} {\cal H}_{\bar{j}}.
\label{eq:sug}
\end{eqnarray}
We denote by ${\cal H}_{i}, {\cal H}_{\bar{i}}$ $(i=1,2,..., n-1)$
the supercurrents associated with
Cartan generators of $sl(n|n-1)$. The $N=2$ super stress tensor
satisfies the following SOPE
\begin{eqnarray}
{{\cal T}}_{sug} (Z_{1}) {{\cal T}}_{sug} (Z_{2})=\frac{c}{z_{12}^{2}} +
\left[ \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} -\frac{\theta_{12}}{z_{12}} {\cal D}+\frac{\bar{\theta}_{12}}{z_{12}} \bar{{\cal D}} +\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \partial \right] {{\cal T}}_{sug}
\label{eq:TT}
\end{eqnarray}
with
\begin{eqnarray} \label{centrN2}
c= -2k\; \alpha_i \alpha_{\bar{j}} g_{i \bar{j}}\;.
\end{eqnarray}
With respect to this ${{\cal T}}_{sug}$, the supercurrents
${\cal H}_{i}, {\cal H}_{\bar{i}} $ are
quasi superprimary superfields of
the dimension $1/2$ with $u(1)$ charge $+1, -1$, respectively.
All other affine $N=2$ supercurrents are superprimary
\begin{eqnarray}
{{\cal T}}_{sug} (Z_{1}) {{\cal J}}_{a (\bar{a})} (Z_{2})=
\left[ s_{a (\bar{a})} \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} -
\frac{\theta_{12}}{z_{12}} {\cal D} + \frac{\bar{\theta}_{12}}{z_{12}} \bar{{\cal D}} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \partial +
q_{a (\bar{a})} \frac{1}{z_{12}} \right] {{\cal J}}_{a (\bar{a})}.
\label{eq:JJ}
\end{eqnarray}
Their dimensions (superspins) $s_{a(\bar{a})}$ and $u(1)$ charges
$q_{a(\bar{a})}$ depend in a certain way
on the parameters $\alpha_{i}, \alpha_{\bar{j}}$. The explicit formulas
for them will be given later, for each specific example
we will consider.
Our last remark in this Section is that for the superalgebra
$sl(n|n-1)$ (and seemingly for other superalgebras admitting a complex
structure) the above $N=2$ extension actually coincides with the
$N=1$ extension given in \cite{DRS,FRS}. In other words, the latter
possesses a {\it hidden}
$N=2$ supersymmetry which becomes manifest in terms
of constrained $N=2$ supercurrents. Indeed, due to the fact that the
generators
of $sl(n|n-1)$ can be divided into the pairs of mutually conjugated ones,
the relevant
$N=1$ supercurrent for each pair is complex and its component content
(two real spins ${1\over 2}$ and two real spins $1$) is such that
these components can be combined into a $N=2$ supermultiplet
\footnote{Similar arguments for the case of bosonic
algebra $g$ were given in \cite{RASS}.}. To explicitly demonstrate
this, let us solve the
constraints (\ref{eq:cons}) via unconstrained
$N=1$ supercurrents $J_{a},J_{\bar{a}}$
\begin{eqnarray}
{{\cal J}}_{a}=J_{a} +\theta^{1} \left[ iD J_{a}+ (-1)^{d_{a}} \frac{1}{k}
{f_{a}}^{bc} J_{b}J_{c} \right] , \nonumber \\
{{\cal J}}_{\bar{a}}=J_{\bar{a}} -\theta^{1} \left[
iD J_{\bar{a}}+ (-1)^{d_{\bar{a}}} \frac{1}{k} {f_{\bar{a}}}^{\bar{b}
\bar{c}}
J_{\bar{b}} J_{\bar{c}} \right].
\label{eq:jj}
\end{eqnarray}
Here the $N=1$ supercurrents are defined on a real $N=1$ superspace
$\widetilde{Z}=(z, \theta^2)$,
$\theta^2 \equiv {1\over 2} (\theta + \bar \theta)$,
$D=\frac{\partial}{\partial {\theta}^{2}} - {\theta}^{2}
\partial$ is a $N=1$ covariant fermionic derivative
and $\theta^{1} \equiv {1\over 2i} (\theta - \bar \theta)$ is an
extra fermionic coordinate.
The SOPEs between the $N=1$ $sl(n|n-1)^{(1)}$ affine supercurrents $J_A (
\widetilde{Z})$
\cite{DRS,FRS}
are given by
\begin{eqnarray}
J_A (\widetilde{Z_1}) J_B (\widetilde{Z_2})=\frac{1}{\widetilde{z_{12}}}
k g_{AB}+\frac{\widetilde{\theta_{12}} }{\widetilde{z_{12}}}
{f_{AB}}^{C} J_C
\label{eq:jj1}
\end{eqnarray}
where
\begin{eqnarray}
\widetilde{\theta_{12}}=\theta_{1}^{2}-\theta_{2}^{2}, \;
\widetilde{z_{12}}=z_{1}-z_{2}-\theta_{1}^{2} \theta_{2}^{2}.
\end{eqnarray}
In a complex basis, the indices $A, B, \cdots $ can be divided into
the two sets of the barred and unbarred indices, thus
demonstrating that the number of $N=1$ supercurrents in the present case
coincides with the number of $N=2$ ones (of course, these complex
$N=1$ supercurrents are reducible, each containing two real $N=1$
supermultiplets). The superalgebra
(\ref{eq:jj1}) is equivalent to the superalgebra ( \ref{eq:qq}) supplemented
with the nonlinear constraints (\ref{eq:cons}). The $N=1$ superfield
formulation
clearly demonstrates that the nonlinearities in the r.h.s. of
eqs. (\ref{eq:qq}) are fake: they appear as the price for manifest
$N=2$ supersymmetry. In what follows the $N=1$ formulation will be
a useful guide of how to impose constraints on the relevant $N=2$
supercurrents corresponding to different embeddings of $sl(2|1)$ into
$sl(n|n-1)$ and to extract those preserving $N=2$ supersymmetry from their
general set.
In the next Sections, we will discuss different hamiltonian reductions
of the $N=2$ $sl(n|n-1)^{(1)}$ current algerbra in $N=2$ superspace for
the particular cases of $n=2, 3$. But before we will sketch the
basic peculiarities of the $N=2$ superspace version of the hamiltonian
reduction procedure.
\section{\bf Hamiltonian Reduction }
\setcounter{equation}{0}
To illustrate the basic idea of different reductions, we start by
considering how we can obtain
extended $N=2$ SCAs by imposing reduction constraints on
the $N=2$ affine supercurrents which we defined in Section $2$.
From now on we will deal with the matrix elements
${\cal J}_{mn}$ of the $sl(n|n-1)$ valued
affine $N=2$ supercurrent (with the $sl(n|n-1)$ generators in the
fundamental representation) rather than with its ajoint
representation components labelled by indices $a, \bar{a}$.
The explicit relation between them is given by
\begin{eqnarray}
{\cal J}_{mn}=
\left( \begin{array}{cccccccccc}
{\cal H}_{\bar{1}}& & & & & & & & \\
& {\cal H}_{\overline{2}}+{\cal H}_{1}& & & & &
\mbox{unbarred} & & \\
& &. & & & & \mbox{indices} & & \\
& & & & & & & & \\
& & & & {\cal H}_{\overline{n-1}}+{\cal H}_{n-2} & & & & \\
& & & & & {\cal H}_{{n-1}} & & & \\
& & & & & & {\cal H}_{\overline{1}}+{\cal H}_{1} & & \\
& \mbox{barred} & & & & & &. & & \\
& \mbox{indices} & & & & & & & \\
& & & & & & & & {\cal H}_{\overline{n-1}}+{\cal H}_{n-1}
\end{array} \right)
\label{eq:jmn}
\end{eqnarray}
(see also Appendices A and B).
We will consider only linear reduction constraints
like in \cite{DS,FRRTW,BTv}.
Then we are led to equate some of ${\cal J}_{mn}$ (we denote the corresponding
subset of indices by the symbol ``hat'') to constants
\begin{eqnarray}
\Phi_{\hat{m} \hat{n}} \equiv {\cal J}_{\hat{m} \hat{n}}-c_{\hat{m} \hat{n}}=0.
\label{eq:constrgen}
\end{eqnarray}
The entries of the constant supermatrix $c_{\hat{m}\hat{n}}$ can be
either $0$, which is
possible both for bosonic and fermionic supercurrents, or $1$, which is
admissible only for bosonic supercurrents. In order to produce $N=2$
supersymmetric algebras these constraints should be invariant with respect
to $N=2$ superconformal transformations generated by improved $N=2$
stress tensor (\ref{eq:sug}), which means that the constrained supercurrents
with nonzero $c_{\hat{m} \hat{n}}$ should have zero spin and
$u(1)$ charge.
In Ref. \cite{FRS}, $W$ superalgebras which can be obtained by the
reductions associated with different embeddings
of $osp(1|2)$ into $sl(n|n-1)$ have been classified in $N=1$ superspace.
Once we know the constraints in $N=1$ superspace, the relation (\ref{eq:jj})
gives us constraints in $N=2$ superspace. Some of the constraints in $N=1$
superspace, being rewritten in $N=2$ superspace, explicitly
break $N=2$ supersymmetry. Meanwhile, we wish to deal with only those
reductions which preserve $N=2$ supersymmetry, because our eventual aim is to
get extended SCAs containing $N=2$ SCA
as a subalgebra. Only a subset of constraints in $N=1$ superspace
preserves $N=2$ supersymmetry, namely those which after substitution
into (\ref{eq:jj}) produce no explicit $\theta$'s in the r.h.s.,
i.e. lead to the $N=2$ constraints in the form (\ref{eq:constrgen}).
Thus, we can choose the approriate subset of constraints in $N=1$
superspace and then extract the constraints in $N=2$
superspace from (\ref{eq:jj}).
Then the first-class constraints, i.e. those which commute among
themselves on the constraints shell, generate a gauge invariance.
An infinitesimal gauge transformation of ${\cal J}_{kl}$ induced by
$\Phi_{\hat{m}\hat{n}}$ with a gauge parameter
${\Lambda}_{\hat{m} \hat{n}}$ can easily be calculated
\begin{eqnarray}
\delta_{\Lambda} {\cal J}_{kl}(Z_{2})=\frac{1}{2\pi i} \oint d
Z_{2} {\Lambda}_{\hat{m} \hat{n}}
(Z_{1}) \left( \Phi_{\hat{m} \hat{n}}(Z_{1}) {\cal J}_{kl}(Z_{2}) \right)
\vert_{ \{\Phi_{\hat{m} \hat{n}}=0\} },
\label{eq:gauge}
\end{eqnarray}
where the symbol $\vert_{ \{ \Phi_{\hat{m} \hat{n}}=0 \} } $ means that
after computing the SOPEs we should pass on the constraints shell by
imposing the constraints (\ref{eq:constrgen}) on the resulting expression
and the gauge parameters $\Lambda_{\hat{m} \hat{n}}$ are general $N=2$
superfields which do not depend on ${\cal J}_{kl}$. It is clear that
the variation of the l.h.s. of (\ref{eq:cons}) vanishes identically
because the SOPEs of (\ref{eq:cons}) with any ${\cal J}_{kl}$, and,
in particular, with $\Phi_{\hat{m} \hat{n}}$ are zero on the shell of
(\ref{eq:cons}) (see discussion in the paragraph below (\ref{eq:DDB})).
{\it By definition}, an extended $N=2$ SCA constructed
by the hamiltonian reduction based on the constraints
(\ref{eq:constrgen})
is a superalgebra generated by gauge invariant
differential - polynomial functionals of affine supercurrents ${\cal J}_{kl}$,
including some $N=2$ stress tensor.
It is possible to find these superalgebras by using
Dirac construction.
Let us remind its main steps.
At first, we should fix the gauge, which means that
we are led to enlarge the original set of first-class constraints by adding
the gauge-fixing conditions (standard gauge-fixing procedure),
such that the total set of constraints becomes second-class.
We denote this extended set of constraints by $\Psi_{\hat{m}\hat{n}}$.
The number of constraints $\Psi_{\hat{m}\hat{n}}$ is
exactly twice the number of $\Phi_{\hat{m}\hat{n}}$.
For the remaining unconstrained supercurrents we will use in this Section
Greek indices, $\alpha, \beta, \cdots$. Clearly, once a gauge freedom with
respect to the $\Lambda$ transformations has been somehow fixed,
the surviving supercurrents ${\cal J}_{\alpha \beta}$ are expressed as
some gauge invariant differential functionals of the original affine
supercurrents.
Secondly, we should construct Dirac brackets between these gauge
invariant supercurrents.
We generalize this procedure to the $N=2$ supersymmetric case and represent
Dirac brackets in an
equivalent form of SOPEs. The new rules for
calculation of SOPEs of the gauge invariant supercurrents which we denote
by brackets with star,
$\left( {\cal J}_{\alpha \beta} (Z_{1}) {\cal J}_{\gamma \sigma} (Z_{2}) \right)^{*}$,
can be defined in terms of original SOPEs of the affine supercurrents
as follows
\vspace{1cm}
\begin{eqnarray}
\left( {\cal J}_{\alpha \beta} (Z_{1}) {\cal J}_{\gamma \sigma} (Z_{2}) \right)^{*}
\equiv
\left( \widetilde{{\cal J}_{
\alpha \beta}} (Z_{1}) \widetilde{{\cal J}_{\gamma \sigma}} (Z_{2}) \right)
\vert_{\{ \Psi_{\hat{m} \hat{n}}=0 \}} \nonumber
\end{eqnarray}
\begin{eqnarray}
- \left[ \frac{1}{(2 \pi i)^2} \oint
\oint d \; Z_{3} d \; Z_{4} \left( \widetilde{{\cal J}_{\alpha \beta}} (Z_{1})
\Psi_{\hat{i} \hat{j}} (Z_{3}) \right) \triangle^{\hat{i} \hat{j},
\hat{k} \hat{l}} (Z_{3},
Z_{4}) \left( \Psi_{\hat{k} \hat{l}} (Z_{4}) \widetilde{{\cal J}_{\gamma \sigma}}
(Z_{2}) \right) \right]
\vert_{\{ \Psi_{\hat{m} \hat{n}}=0 \}},
\label{eq:dirac}
\end{eqnarray}
where $\widetilde{{\cal J}_{\alpha \beta}}$ are functionals of the original
supercurrents ${\cal J}_{kl}$ (including both unconstrained
${\cal J}_{\alpha \beta}$ and constrained ${\cal J}_{\hat{m} \hat{n}}$
supercurrents) which satisfy the following restrictions on the constraints
shell
\begin{eqnarray}
\widetilde{{\cal J}_{\alpha \beta}} \vert_{\{ \Psi_{\hat{m}\hat{n}}=0 \}}=
{\cal J}_{\alpha \beta}
\label{eq:tilj}
\end{eqnarray}
and are arbitrary otherwise, the supermatrix
$\triangle^{\hat{i} \hat{j},\hat{k} \hat{l}}(Z_{1},Z_{2})$ is the
inverse of the supermatrix
\begin{eqnarray}
\triangle_{\hat{i} \hat{j}, \hat{k} \hat{l}}(Z_{1},Z_{2})=
\left( \Psi_{\hat{i} \hat{j}} (Z_{1}) \Psi_{\hat{k} \hat{l}} (Z_{2}) \right)
\vert_{\{ \Psi_{\hat{m} \hat{n}}=0 \}}
\end{eqnarray}
i.e.
\begin{eqnarray}
\frac{1}{2 \pi i} \oint d \; Z_{2} \triangle^{\hat{i} \hat{j}, \hat{k}
\hat{l}} (Z_{1}, Z_{2})
\triangle_{
\hat{k} \hat{l}, \hat{m} \hat{n}} (Z_{2}, Z_{3})= \delta^{\hat{i}}_{\hat{m}}
\delta^{\hat{j}}_{\hat{n}}
\theta_{13} {\bar{\theta}}_{13}
\delta(z_{1}-z_{3}) \;.
\end{eqnarray}
Any gauge invariant supercurrent can be represented as some functional of
${\cal J}_{\alpha \beta}$ and SOPEs between these functionals can be
calculated using SOPEs
(\ref{eq:dirac}) between ${\cal J}_{\alpha \beta}$.
It is a very complicated technical problem to calculate the inverse
supermatrix $\triangle^{\hat{i} \hat{j},\hat{k} \hat{l}}(Z_{1},Z_{2})$ in
the general case. To get round this difficulty,
we use the following trick. By looking at
(\ref{eq:dirac}), one can observe that for $\widetilde{{\cal J}_{\alpha \beta}}$
satisfying
\begin{eqnarray}
\left( \widetilde{{\cal J}_{\alpha \beta}} (Z_{1}) \Psi_{\hat{m} \hat{n}} (Z_{2})
\right) \vert_{\{ \Psi_{\hat{m} \hat{n}}=0 \}} = 0
\label{eq:req}
\end{eqnarray}
the second term in the r.h.s. of (\ref{eq:dirac}) is vanishing.
We are at freedom to choose $\widetilde{{\cal J}_{\alpha \beta}}$ to satisfy
eq.(\ref{eq:req}) as these functionals are {\it a priori} arbitrary up to
the condition (\ref{eq:tilj}) which is obviously consistent
with (\ref{eq:req}). Then
the SOPEs with star between the gauge invariant supercurrents
coincide with ordinary SOPEs between
$\widetilde{{\cal J}_{\alpha \beta}}$ on the constraints shell and so
can be calculated using SOPEs (\ref{eq:qq}) for the original
affine supercurrents
\begin{eqnarray}
\left( {\cal J}_{\alpha \beta} (Z_{1}) {\cal J}_{\gamma \sigma} (Z_{2}) \right)^{*}
& \equiv &
\left( \widetilde{{\cal J}_{
\alpha \beta}} (Z_{1}) \widetilde{{\cal J}_{\gamma \sigma}} (Z_{2})
\right) \vert_{\{ \Psi_{\hat{m} \hat{n}}=0 \}}.
\label{eq:req1}
\end{eqnarray}
In this way, the task of constructing $N=2$ extended superalgebras reduces
to that of constructing the functionals
$\widetilde{{\cal J}_{\alpha \beta}}$ satisfying
the restrictions (\ref{eq:tilj}), (\ref{eq:req}).
Now let us discuss the general structure of such functionals.
It is evident that only those of them which are {\it linear} in the
total set of constraints $\Psi_{\hat{m} \hat{n}}$ can actually
contribute to (\ref{eq:req1}), because the SOPEs including any higher order
monomial of $\Psi_{\hat{m} \hat{n}}$ are proportional to
$\Psi_{\hat{m} \hat{n}}$ and so obviously vanish on the constraints shell
$\{ \Psi_{\hat{m} \hat{n}} = 0 \}$. The coefficients in these linear
functionals can in general be nonlinear functionals of the remaining
unconstrained supercurrents ${\cal J}_{\alpha \beta}$.
Keeping this in mind, from now on we consider as a starting expression for
$\widetilde{{\cal J}_{\alpha \beta}}$
linear functionals of constraints $\Psi_{\hat{m} \hat{n}}$
(and derivatives of the latter) with nonlinear
in general coefficient-functions of ${\cal J}_{\alpha \beta}$. Taking for these
coefficients the
most general ansatz in terms of ${\cal J}_{\alpha \beta}$ with arbitrary
constant coefficients, such that it preserves superspins and $u(1)$ charges
with respect to the improved $N=2$ stress tensor (\ref{eq:sug}), and
substituting it into eqs.
(\ref{eq:tilj}), (\ref{eq:req}), one obtains the solution which proves
to be unique up to some unessential
coefficients which do not contribute to (\ref{eq:req1}).
In the next Section we will illustrate the formalism described above
by the simplest example of hamiltonian reduction of the
$N=2$ $sl(2|1)^{(1)}$ superalgebra.
\section{\bf Example: $N=2$ $sl(2|1)^{(1)}$ Affine Superalgebra}
\setcounter{equation}{0}
Let us apply the general procedure developed in the previous Section to
the superalgebra $N=2$ $sl(2|1)^{(1)}$. We will naturally come to the
$N=2$ superspace formulation of standard $N=2$ SCA in this way.
In Appendix A, for completeness we give the explicit form of generators,
structure constants and Killing metric for $sl(2|1)$ superalgebra in
the complex basis described in Section $2$, as well as the relations
between affine supercurrents ${\cal J}_{a}, {\cal J}_{\bar{a}}$ in this basis and
matrix elements ${\cal J}_{mn}$ introduced in Section $3$.
Substituting these formulas into (\ref{eq:qq}), (\ref{eq:cons}) and
(\ref{eq:sug}) one can obtain explicit expressions for the defining SOPEs
of $N=2$ affine extension of $sl(2|1)$,
for nonlinear constraints the relevant supercurrents satisfy, as well as
for the improved Sugawara $N=2$ stress tensor.
The last one has the following form
\begin{eqnarray}
{\cal T}_{sug}=\frac{1}{k} \left( {\cal J}_{12} {\cal J}_{21} - {\cal H}_{1} {\cal H}_{\bar{1}} -
{\cal J}_{13} {\cal J}_{31}-{\cal J}_{23} {\cal J}_{32} \right) +
\alpha_{1} \bar{{\cal D}} {\cal H}_{1}+\alpha_{\bar{1}}
{\cal D} {\cal H}_{\bar{1}},
\label{eq:sugsl21}
\end{eqnarray}
where two parameters, $\alpha_{1}$ and $\alpha_{\bar{1}}$, give rise to a
splitting of supercurrents into the grades with positive, zero and negative
dimensions and $u(1)$ charges (see Table 1).
Actually, this splitting is due to the existence of two
grading operators: $(\alpha_{1} t_{2} +\alpha_{\bar{1}} t_{\bar{2}} )/2 $ and
$\alpha_{1} t_{2}-\alpha_{\bar{1}} t_{\bar{2}}$, $t_i, t_{\bar{i}}$ being
Cartan generators of $sl(2|1)$ in the coadjoint representation.
The eigenvalues of the former are exactly
( ``dimension''- $1/2$) in the Table 1 and those of the latter are
(``$u(1)$ charge'' $\pm 1$) where $+1$ is for barred supercurrents
and $-1$ for unbarred ones.
\begin{eqnarray}
\begin{array}{ccc}
\mbox{Table 1} \nonumber \\ \hline
\mbox{scs} & \mbox{dim} & u(1) \nonumber \\
\hline
\vspace{2mm}
{\cal H}_{1}^F & 1/2 & 1 \nonumber \\
\vspace{2mm}
{\cal H}_{\bar{1}}^F & 1/2 & -1 \nonumber \\
\vspace{2mm}
{\cal J}_{12}^F & (1+\alpha_{1}-\alpha_{\bar{1}})/2 &
(1+\alpha_{1}+\alpha_{\bar{1}}) \nonumber \\
\vspace{2mm}
{\cal J}_{21}^F & (1-\alpha_{1}+\alpha_{\bar{1}})/2 &
(-1-\alpha_{1}-\alpha_{\bar{1}}) \nonumber \\
\vspace{2mm}
{\cal J}_{13}^B & (1+\alpha_{1})/2 & (1+\alpha_{1}) \nonumber \\
\vspace{2mm}
{\cal J}_{31}^B & (1-\alpha_{1})/2 & (-1-\alpha_{1}) \nonumber \\
\vspace{2mm}
{\cal J}_{23}^B & (1+\alpha_{\bar{1}})/2 & (1-\alpha_{\bar{1}}) \nonumber \\
\vspace{2mm}
{\cal J}_{32}^B & (1-\alpha_{\bar{1}})/2 & (-1+\alpha_{\bar{1}}) \nonumber \\
\hline
\end{array}
\end{eqnarray}
In this and all subsequent Tables we use the following abbreviations:
``scs'' for supercurrents, ``dim'' for superconformal dimensions
and ``$u(1)$''
for $u(1)$ charges.
We also give the explicit form of the nonlinear constraints (\ref{eq:cons})
\begin{eqnarray}
{\cal D} {\cal H}_{1} = 0, & \;\; & \bar{{\cal D}} {\cal H}_{\bar{1}} = 0, \nonumber \\
\left( {\cal D} + \frac{1}{k} {\cal H}_{1} \right) {\cal J}_{12} = 0, & \;\; &
\left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{1}} \right) {\cal J}_{21} = 0, \nonumber \\
{\cal D} {\cal J}_{13} - \frac{1}{k} {\cal J}_{12} {\cal J}_{23} = 0, & \;\; &
\left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{1}} \right) {\cal J}_{31} - \frac{1}{k}
{\cal J}_{21} {\cal J}_{32} = 0, \nonumber \\
\left( {\cal D} - \frac{1}{k} {\cal H}_{1} \right) {\cal J}_{23} = 0, & \;\; &
\bar{{\cal D}} {\cal J}_{32} = 0.
\label{eq:constra}
\end{eqnarray}
Now we are ready to consider a hamiltonian reduction of $N=2$
$ sl(2|1)^{(1)}$
which produces $N=2$ SCA. To this end, we should first learn at which
values of parameters $\alpha_{1}$ and $\alpha_{\bar{1}}$ at least one
of the bosonic supercurrents could have the spin
and $u(1)$ charge characteristic of the $N=2$ stress tensor, i.e.
$1, 0$, respectively. It turns out possible with the following choice
\begin{eqnarray}
\alpha_{1} = -1, \;\; \alpha_{\bar{1}} = 1.
\label{eq:alphab}
\end{eqnarray}
In this case, besides the fermionic supercurrents ${\cal H}_{1},
{\cal H}_{\bar{1}}$ with
the spin and $u(1)$ charge $1/2$ and $\pm 1$,
$N=2$ $ sl(2|1)^{(1)}$ superalgebra contains bosonic spin $0$
$( {\cal J}_{13}, {\cal J}_{32}) $ and spin $1$ $({\cal J}_{31}, {\cal J}_{23} )$
ones with zero $u(1)$ charges, as well as the fermionic doublet
${\cal J}_{12}, {\cal J}_{21}$ with spins $-1/2,3/2$ and $u(1)$ charges $1, -1$,
respectively.
Secondly, we should put first-class constraints on some supercurrents at
which at least one of two spin $1$ supercurrents
( ${\cal J}_{31} \; \mbox{or} \; {\cal J}_{23}$ )
is unconstrained in order to be able to identify it with
$N=2$ unconstrained stress tensor. At first sight, it seems
impossible to achieve this because from the
beginning all the supercurrents are constrained by the
conditions (\ref{eq:constra}). Nevertheless, it can be done.
Let us briefly explain the basic idea of how unconstrained $N=2$
superfields can come out in this way.
By looking at the constraints (\ref{eq:constra}), one sees that they are
quadratically nonlinear and their number precisely matches with that of
supercurrrents. Moreover, in every constraint there is only one linear term
with spinor covariant derivative on some supercurrent, and different
constraints contain different linear terms, so they are in one-to-one
correspondence with the consistent set of standard chiral and
anti-chiral conditions. The last ones reduce the number of
independent superfield components by the factor two.
The same is evidently true for a nonlinear generalization of
these constraints (\ref{eq:constra}): the only new point is that
the components which were forced to be zero in
the case of chiral constraints become some functions of the remaining
independent ones in the case of (\ref{eq:constra}). However, an
important difference of the latter from the linear constraints
is the following. If we replace
some bosonic supercurrents in (\ref{eq:constra}) by nonzero constants,
then in some constraints the nonlinear terms can produce {\it a linear} one
without a spinor derivative on it. So, this constraint becomes algebraic
with respect to the supercurrent entering it linearly and can be solved
for the latter. Thus this supercurrent turns out to be eventually expressed
in terms of other ones and their spinor covariant derivatives. Now among
the remaining independent supercurrents one can find, in a number of cases,
unconstrained $N=2$ superfields. This is just what comes about in the
case at hand. An analogous resume
could be drawn from the analysis of solutions of $N=2$ constraints
(\ref{eq:constra}) in terms of unconstrained $N=1$ superfields (\ref{eq:jj}).
Keeping in mind the above remark, we choose first-class constraints as
follows
\begin{eqnarray}
{\cal J}_{mn}^{constr}=
\left( \begin{array}{ccc}
\ast & 0 & 1 \\
\ast & \ast & \ast \\
\ast & 1 & \ast
\end{array} \right).
\label{eq:sl21constr}
\end{eqnarray}
They clearly preserve $N=2$ superconformal symmetry generated by
${\cal T}_{sug}$ (\ref{eq:sugsl21}), (\ref{eq:alphab}).
This set of constraints is also consistent
with eqs. (\ref{eq:constra}). Indeed, by substituting (\ref{eq:sl21constr})
into (\ref{eq:constra}) we find that those constraints from (\ref{eq:constra})
which include spinor derivative of the supercurrents
${\cal J}_{13}, {\cal J}_{32}$ and ${\cal J}_{12}$ are satisfied identically while the
constraint containing spinor derivative of ${\cal J}_{31}$ current becomes
algebraic and expresses ${\cal J}_{21}$ in terms of ${\cal J}_{31}$
\begin{eqnarray}
{\cal J}_{21} = k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{1}} \right) {\cal J}_{31}.
\label{eq:constsol}
\end{eqnarray}
The remaining constraints from the set (\ref{eq:constra}) preserve
their form on the constraints
shell (\ref{eq:sl21constr}). Thus on the shell of
constraints (\ref{eq:sl21constr}) there arise no any restrictions on the
spin $1$, $u(1)$ charge $0$ bosonic supercurrent ${\cal J}_{31}$, so
the latter is an
unconstrained $N=2$ superfield and, as we will see soon, proves to be
directly related to the $N=2$ superconformal stress tensor.
Let us note that the constraints (\ref{eq:sl21constr}) actually amount
to the set of constraints imposed in \cite{DRS}
in $N=1$ superspace. This latter set can be shown to
produce the above
constraints without breaking $N=2$ supersymmetry through the explicit
relation (\ref{eq:jj}) between $N=1$ and $N=2$ supercurrents.
Constraints (\ref{eq:sl21constr}) can easily be checked to have zero mutual
SOPEs on their shell, so they are first-class and give rise to a gauge
invariance which can be used to gauge away three more entries in the
supermatrix (\ref{eq:sl21constr}). Indeed, with respect to infinitesimal
gauge transformations
(\ref{eq:gauge}) generated by constraints (\ref{eq:sl21constr}) with the gauge
parameters $\Lambda_{12}, \Lambda_{13}$ and $\Lambda_{32}$ the currents
${\cal J}_{23}, {\cal H}_{1}$ and ${\cal H}_{\bar{1}}$ are transformed inhomogeneously
\begin{eqnarray}
\delta_{\Lambda_{12}} {\cal J}_{23} (Z_{2}) & = & -
\left( {\cal D} -\frac{1}{k} {\cal H}_{1} \right)\Lambda_{12},\nonumber\\
\delta_{\Lambda_{13}} {\cal H}_{1} (Z_{2}) & = & {\cal D} {\Lambda}_{13}, \nonumber \\
\delta_{\Lambda_{32}} {\cal H}_{\bar{1}} (Z_{2}) & = & \bar{{\cal D}} {\Lambda}_{32}.
\end{eqnarray}
One can explicitly check that these gauge transformations preserve
the constraints (\ref{eq:constra}).
As a result, we can consistently fix the gauge as\footnote{In this gauge
there remains a residual gauge freedom with chiral $\Lambda_{12}$ and
anti-chiral $\Lambda_{13}$, $\Lambda_{32}$. However, the final expression
for $N=2$ stress tensor turns out to be invariant under this
residual freedom.}
\begin{eqnarray}
{\cal J}_{23} = 0, \;\; {\cal H}_{1} = 0, \;\; {\cal H}_{\bar{1}} = 0.
\label{eq:gagfix}
\end{eqnarray}
It is easy to check that the total set of constraints, i.e. constraints
(\ref{eq:sl21constr}) and gauge fixing conditions (\ref{eq:gagfix}),
is second-class.
Substituting the gauge fixing conditions (\ref{eq:gagfix}) and
the expression (\ref{eq:constsol}) into supermatrix (\ref{eq:sl21constr})
we obtain the expression
for $N=2$ supercurrents ${\cal J}_{mn}$ in the highest
weight ( or Drinfeld-Sokolov \cite{DS} ) gauge
\begin{eqnarray}
{\cal J}_{mn}^{DS}=
\left( \begin{array}{ccc}
0 & 0 & 1 \\
k \bar{{\cal D}} {\cal J}_{31} & 0 & 0 \\
{\cal J}_{31} & 1 & 0
\end{array} \right).
\label{eq:ds}
\end{eqnarray}
So the superalgebra which is produced from $N=2$ $sl(2|1)^{(1)}$ by
hamiltonian reduction associated with the constraints (\ref{eq:sl21constr})
is generated by only one gauge invariant bosonic supercurrent
$\widetilde{{\cal J}_{31}}$
which coincides with ${\cal J}_{31}$ on the shell of total set of constraints
(see (\ref{eq:tilj})).
Our next task is to find $\widetilde{{\cal J}_{31}}$ from the
conditions (\ref{eq:tilj}), (\ref{eq:req}).
This can be easily done by making use of the general procedure described
in Section 3. As a result we obtain the following expression for
$\widetilde{{\cal J}_{31}}$ up to unessential terms
\begin{eqnarray}
\widetilde{{\cal J}_{31}}=\left( -{\cal J}_{12} {\cal J}_{21}+{\cal J}_{13}
{\cal J}_{31}+{\cal J}_{23} \right)+k \bar{{\cal D}} {\cal H}_{1}- k {\cal D} {\cal H}_{\bar{1}}+k^3
\bar{{\cal D}} {\cal J}_{12}'+k^2 {\cal J}_{13}'-k^2 {\cal J}_{32}'.
\end{eqnarray}
Substituting this expression into (\ref{eq:req1}), we get the
SOPE of superalgebra we are looking for. This SOPE coincides with
SOPE of $N=2$
SCA (\ref{eq:TT}) with central charge $-2 k$ after rescaling
\begin{eqnarray}
{\cal J}_{31} \rightarrow -k {\cal J}_{31}.
\end{eqnarray}
Before closing this Section, we briefly mention that there
exists another choice for the gauge fixing, so called
diagonal gauge \cite{DS}. Namely,
\begin{eqnarray}
{\cal J}_{23} = 0, \;\; {\cal J}_{21} = 0, \;\; {\cal J}_{31} = 0.
\label{eq:gagfix1}
\end{eqnarray}
Repeating all the steps we have passed before, one can obtain the
following form for $N=2$ supercurrents ${\cal J}_{mn}$ in this case
\begin{eqnarray}
{\cal J}_{mn}^{diag}=
\left( \begin{array}{ccc}
{\cal H}_{\bar{1}} & 0 & 1 \\
0 & {\cal H}_{1} & 0 \\
0 & 1 & {\cal H}_{\bar{1}}+{\cal H}_{1}
\end{array} \right),
\label{eq:dia}
\end{eqnarray}
where ${\cal H}_{1}, {\cal H}_{\bar{1}}$ are chiral and anti-chiral $N=2$ fermionic
superfields which form the $N=2$ $u(1)$ affine superalgebra
\begin{eqnarray}
{\cal H}_{1} (Z_{1}) {\cal H}_{\bar{1}} (Z_{2}) = \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \frac{k}{2} - \frac{1}{z_{12}} k.
\end{eqnarray}
Two different gauge choices (\ref{eq:ds}), (\ref{eq:dia}) are
connected to each other by some gauge
transformation. If we would know this gauge transformation, then we could
obtain the standard Miura free field realization of $N=2$ SCA in terms of
chiral and
anti-chiral fermionic superfields. However, in our simple case it is
of no need to know this gauge transformation for deducing
Miura realization, if we observe that the total set of constraints
(\ref{eq:sl21constr}), (\ref{eq:gagfix1}) for the diagonal gauge on
the constraint shell is
invariant under transformations generated by the $N=2$ stress tensor
${\cal T}_{sug}$ (\ref{eq:sugsl21}), (\ref{eq:alphab}).
This means that ${\cal T}_{sug}$ is gauge invariant and on the constraints shell
has the following form
\begin{eqnarray}
{\cal T}_{sug}= - \frac{1}{k} {\cal H}_{1} {\cal H}_{\bar{1}} -
\bar{{\cal D}} {\cal H}_{1} + {\cal D} {\cal H}_{\bar{1}}
\label{eq:sugsl21shell}
\end{eqnarray}
which coincides with the standard Miura free field form of
the $N=2$ stress tensor.
In the next Section we will discuss various reductions of
$N=2$ $sl(3|2)^{(1)}$ and deduce some new superfield extended $N=2$
SCAs in this way.
\section{\bf Hamiltonian Reductions of
$N=2$ $sl(3|2)^{(1)}$ Affine Superalgebra}
\setcounter{equation}{0}
The reductions of $N=2$ $sl(3|2)^{(1)}$ we will consider in this Section
give rise to four new types of extensions of $N=2$ SCA.
The first one is rather unusual in the sense that
the $N=2$ stress tensor is a constrained supercurrent. The second possesses
an unconstrained stress tensor, but contains spin $0$ supercurrents, such
that it turns out impossible to
decouple dimension $0$ component currents. We will concentrate
on the third and fourth cases corresponding to $N=2$ $u(2|1)$ and
$N=2$ $u(3)$ SCAs, respectively, because these are ``canonical'' in the sense
that the relevant $N=2$ stress tensor is unconstrained and there are
no spin $0$ supercurrents. We will also illustrate
how the known $N=2$ $W_3$ \cite{LPRSW} and $N=2$ $W_{3}^{(2)}$ \cite{IKS}
SCAs reappear in the hamiltonian reduction approach in $N=2$ superspace.
It is rather straightforward to find the structure constants and Killing
metric in the complex basis for $sl(3|2)$, so we do not write down them
explicitly (see Appendix B). From the general expression for the
improved Sugawara $N=2$ stress tensor
(\ref{eq:sug}) we obtain it for $N=2$ $sl(3|2)^{(1)}$ in the following form
\begin{eqnarray}
{\cal T}_{sug} & = &
\frac{1}{k} \left( -{\cal H}_{1} {\cal H}_{\bar{1}}-{\cal H}_{2} {\cal H}_{\bar{2}}+
{\cal J}_{12} {\cal J}_{21}+{\cal J}_{13} {\cal J}_{31}+{\cal J}_{23} {\cal J}_{32}-{\cal J}_{14} {\cal J}_{41}
\right. \nonumber \\
& & \left.
-{\cal J}_{15} {\cal J}_{51}-{\cal J}_{24} {\cal J}_{42}-{\cal J}_{25} {\cal J}_{52}-{\cal J}_{34} {\cal J}_{43}-
{\cal J}_{35} {\cal J}_{53}-{\cal J}_{45} {\cal J}_{54}-{\cal H}_{2} {\cal H}_{\bar{1}} \right) \nonumber \\
& & +\alpha_{1} \bar{{\cal D}} {\cal H}_{1}+\alpha_{\bar{1}} {\cal D} {\cal H}_{\bar{1}}+\alpha_{2}
\bar{{\cal D}} {\cal H}_{2}+\alpha_{\bar{2}} {\cal D} {\cal H}_{\bar{2}},
\label{eq:sugsl32}
\end{eqnarray}
where four parameters, $\alpha_{1}, \alpha_{\bar{1}}, \alpha_{2},
\alpha_{\bar{2}}$ split the supercurrents into the grades with positive,
zero and negative dimensions and $u(1)$ charges (see Table 2).
Let us stress that the nonlinear constraints (\ref{eq:cons}) for the case
of $sl(3|2)$ can be easily read off using the structure constants
of this superalgebra. They will play the important role in all the
calculations in the remainder of this paper.
Our main aim in this Section will be to find
extended $N=2$ SCAs which contain at least one
{\it bosonic unconstrained} supercurrent with dimension $1$ and vanishing
$u(1)$ charge by applying the general procedure of the hamiltonian
reduction to $N=2$ $sl(3|2)^{(1)}$ affine superalgebra.
In the next Subsections, we will present only the basic results and make
some comments without detailed explanations, because most
of technical points are a direct generalization of those
expounded in Section $4$ on the simpler example of $N=2$ $sl(2|1)^{(1)}$.
\begin{eqnarray}
\begin{array}{ccc}
\mbox{Table 2} \nonumber \\
\hline
\mbox{scs} & \mbox{dim} & u(1) \nonumber \\ \hline
\vspace{2mm}
{\cal H}_{\bar{1}}^F & 1/2 & -1 \nonumber \\
\vspace{2mm}
{\cal H}_{1}^F & 1/2 & 1 \nonumber \\
\vspace{2mm}
{\cal H}_{\bar{2}}^F & 1/2 & -1 \nonumber \\
\vspace{2mm}
{\cal H}_{2}^F & 1/2 & 1 \nonumber \\
\vspace{2mm}
{\cal J}_{12}^F & (1+\alpha_{1}-\alpha_{\bar{1}}+\alpha_{\bar{2}})/2 &
(1+\alpha_{1}+\alpha_{\bar{1}}-
\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{21}^F & (1-\alpha_{1}+\alpha_{\bar{1}}-\alpha_{\bar{2}})/2 &
(-1-\alpha_{1}-\alpha_{\bar{1}}+
\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{13}^F & (1-\alpha_{\bar{1}}+\alpha_{2})/2 & (1+\alpha_{\bar{1}}+
\alpha_{2}) \nonumber \\
\vspace{2mm}
{\cal J}_{31}^F & (1+\alpha_{\bar{1}}-\alpha_{2})/2 & (-1-\alpha_{\bar{1}}-
\alpha_{2}) \nonumber \\
\vspace{2mm}
{\cal J}_{23}^F & (1-\alpha_{1}+\alpha_{2}-\alpha_{\bar{2}})/2 &
(1-\alpha_{1}+\alpha_{2}+
\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{32}^F & (1+\alpha_{1}-\alpha_{2}+\alpha_{\bar{2}})/2 &
(-1+\alpha_{1}-
\alpha_{2}-\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{14}^B & (1+\alpha_{1})/2 & (1+\alpha_{1}) \nonumber \\
\vspace{2mm}
{\cal J}_{41}^B & (1-\alpha_{1})/2 & (-1-\alpha_{1}) \nonumber \\
\vspace{2mm}
{\cal J}_{15}^B & (1-\alpha_{\bar{1}}+\alpha_{2}+\alpha_{\bar{2}})/2 &
(1+\alpha_{\bar{1}}+\alpha_{2}-
\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{51}^B & (1+\alpha_{\bar{1}}-\alpha_{2}-\alpha_{\bar{2}})/2 &
(-1-\alpha_{\bar{1}}-\alpha_{2}+
\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{24}^B & (1+\alpha_{\bar{1}}-\alpha_{\bar{2}})/2 & (1-\alpha_{\bar{1}}+
\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{42}^B & (1-\alpha_{\bar{1}}+\alpha_{\bar{2}})/2 & (-1+\alpha_{\bar{1}}-
\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{25}^B & (1-\alpha_{1}+\alpha_{2})/2 & (1-\alpha_{1}+\alpha_{2}) \nonumber \\
\vspace{2mm}
{\cal J}_{52}^B & (1+\alpha_{1}-\alpha_{2})/2 & (-1+\alpha_{1}-\alpha_{2}) \nonumber \\
\vspace{2mm}
{\cal J}_{34}^B & (1+\alpha_{1}+\alpha_{\bar{1}}-\alpha_{2})/2 & (1+\alpha_{1}-
\alpha_{\bar{1}}-\alpha_{2}) \nonumber \\
\vspace{2mm}
{\cal J}_{43}^B & (1-\alpha_{1}-\alpha_{\bar{1}}+\alpha_{2})/2 & (-1-\alpha_{1}+
\alpha_{\bar{1}}+
\alpha_{2}) \nonumber \\
\vspace{2mm}
{\cal J}_{35}^B & (1+\alpha_{\bar{2}})/2 & (1-\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{53}^B & (1-\alpha_{\bar{2}})/2 & (-1+\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{45}^F & (1-\alpha_{1}-\alpha_{\bar{1}}+\alpha_{2}+\alpha_{\bar{2}})/2 &
(1-\alpha_{1}+
\alpha_{\bar{1}}+\alpha_{2}-\alpha_{\bar{2}}) \nonumber \\
\vspace{2mm}
{\cal J}_{54}^F & (1+\alpha_{1}+\alpha_{\bar{1}}-\alpha_{2}-\alpha_{\bar{2}})/2
& (-1+\alpha_{1}-
\alpha_{\bar{1}}-\alpha_{2}+\alpha_{\bar{2}}) \nonumber \\
\hline
\end{array}
\end{eqnarray}
\subsection{\bf $N=2$ $W_{3}$ SCA}
In order to understand the reduction scheme in the case under consideration,
we take as a first example $N=2$ $ W_{3}$
SCA \cite{LPRSW} and study how it is reproduced in our method.
The algebra $N=2$ $W_{3}$ has one extra spin $2$ bosonic
supercurrent besides the spin $1$ $N=2$ stress tensor. This counting
suggests that we should impose ten constraints which is the ``maximal''
set. The point is that requiring the constraints to be first-class
restricts a possible number of such constraints. It can be
easily checked that this requirement cannot be met
if the number of constraints exceeds ten.
For the choice
\begin{eqnarray}
\alpha_{1}=-1, \;\; \alpha_{\bar{1}}=2, \;\; \alpha_{2}=-2,
\;\; \alpha_{\bar{2}}=1
\label{eq:w3para}
\end{eqnarray}
in Table 3 we give the list of ``twisted'' dimensions and $u(1)$ charges of
those supercurrents which will be subjected to the reduction constraints
and corresponding gauge fixing conditions (we use for them, respectively,
the abbreviation ``constr. scs'' and ``g.f. scs'').
\begin{eqnarray}
\begin{array}{ccccccccccc}
\mbox{Table 3} \nonumber \\
\hline
u(1) & 1 & 0 & 0 & 1 & 1 & 1 & 0 & 0 & 0 & 0 \nonumber \\
\mbox{dim} & -\frac{3}{2} & -1 & -1 & -\frac{1}{2} & -\frac{1}{2} &
-\frac{1}{2} & 0 & 0 & 0 & 0 \nonumber \\
\hline
\mbox{constr. scs} & {\cal J}_{13}^F & {\cal J}_{15}^B & {\cal J}_{43}^F &
{\cal J}_{12}^F & {\cal J}_{23}^F &
{\cal J}_{45}^F & {\cal J}_{14}^B & {\cal J}_{25}^B & {\cal J}_{42}^F & {\cal J}_{53}^B \nonumber \\
\hline
\hline
\mbox{g.f. scs} & {\cal J}_{34}^B &{\cal J}_{52}^B & {\cal J}_{32}^F &
{\cal J}_{24}^B &
{\cal J}_{35}^B & {\cal J}_{54}^F & {\cal H}_{1}^F & {\cal H}_{2}^F & {\cal H}_{\bar{1}}^F &
{\cal H}_{\bar{2}}^F \nonumber \\ \hline
\mbox{dim} & 2 & 1 & \frac{3}{2} & 1 & 1 & \frac{3}{2} & \frac{1}{2}
& \frac{1}{2} & \frac{1}{2} & \frac{1}{2} \nonumber \\
u(1) & 0 & 0 & -1 &
0 & 0 & -1 & 1 & 1 & -1 & -1 \nonumber \\
\hline
\end{array}
\end{eqnarray}
We impose the constraints on all the negative and zero dimension
supercurrents as is summarized below
\begin{eqnarray}
{\cal J}_{mn}^{constr}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & 0 \\
\ast & \ast & 0 & \ast & 1 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & 1 & \ast & \ast
\end{array} \right).
\label{eq:w3cons}
\end{eqnarray}
As was repeatedly mentioned above, these first-class constraints generate
gauge invariances. In the upper line of Table 3 we place the supercurrents
which are subjected to the above constraints and are basically the
generators of these invariances according to the general formula
(\ref{eq:gauge}). The lower line collects the supercurrents which are
gauged away by these invariances. For example,
${\cal J}_{34}$ can be
gauged away using the gauge transformation generated by
constraint ${\cal J}_{13}$ (${\cal J}_{52}$ by ${\cal J}_{15}$ and so on).
Note that four constraints of units in (\ref{eq:w3cons}) are necessary
to gauge away four dimension $1/2$ supercurrents
corresponding to Cartan elements.
As we see, only four supercurrents ${\cal J}_{21}, {\cal J}_{31}, {\cal J}_{41}, {\cal J}_{51}$
eventually survive. Substituting (\ref{eq:w3cons})
into the nonlinear constraints (\ref{eq:cons}) we
find that ${\cal J}_{21}, {\cal J}_{31}$, before fixing the gauge, are expressed
as follows
\begin{eqnarray}
{\cal J}_{21} & = & k \left( \bar{{\cal D}}-\frac{1}{k} {\cal H}_{\bar{1}} \right) {\cal J}_{41}, \nonumber \\
{\cal J}_{31} & = & k \left( \bar{{\cal D}}-\frac{1}{k} \left( {\cal H}_{\bar{1}}+{\cal H}_{\bar{2}}
\right) \right) {\cal J}_{51}- {\cal J}_{21} {\cal J}_{52}- {\cal J}_{41} {\cal J}_{54}.
\label{eq:linear}
\end{eqnarray}
After gauging away the unphysical degrees of
freedom in accord with Table 3, we are left with the following supercurrent
matrix ${\cal J}_{mn}$ in the highest weight gauge
\begin{eqnarray}
{\cal J}_{mn}^{DS}=
\left( \begin{array}{ccccc}
0 & 0 & 0 & 1 & 0 \\
k \bar{{\cal D}} {\cal J}_{41} & 0 & 0 & 0 & 1 \\
k \bar{{\cal D}} {\cal J}_{51} & 0 & 0 & 0 & 0 \\
{\cal J}_{41} & 1 & 0 & 0 & 0 \\
{\cal J}_{51} & 0 & 1 & 0 & 0
\end{array} \right),
\label{eq:w3ds}
\end{eqnarray}
Thus as an output we have two independent unconstrained supercurrents with
zero $u(1)$ charges: a dimension $1$ supercurrent ${\cal J}_{41}$
which is nothing but the $N=2$ stress tensor and a
dimension $2$ supercurrent ${\cal J}_{51}$.
We will not discuss here how to construct gauge invariant supercurrents
and which SOPEs they satisfy, because all these
formulas can be reproduced via a secondary
hamiltonian reduction from $N=2$ $W_{3}^{(2)}$ SCA which will be
discussed in the following
Subsection. Anticipating the result, the relevant set of SOPEs forms the
classical $N=2$ $W_3$ SCA \cite{LPRSW}.
\subsection{\bf $N=2$ $W_{3}^{(2)}$ SCA}
Let us now describe another reduction.
We wish to understand how $N=2$ $W_{3}^{(2)}$ SCA of Ref. \cite{IKS}
can be obtained within our procedure. Recall that
this algebra is described in $N=2$ superspace by
the spin $1/2, 2$ bosonic and $1/2, 2$ fermionic constrained supercurrents
in addition to the spin 1 bosonic unconstrained $N=2$ stress tensor.
To match this superfield content, we are led to impose nine constraints on
the $N=2$ affine supercurrents. One could try to proceed by relaxing one of
the constraints (\ref{eq:w3cons}), still with the same choice
of the splitting parameters (\ref{eq:w3para}).
However, in this basis one finds no spin 2 fermionic supercurrents required
by the superfield content of $N=2$ $W_3^{(2)}$ SCA.
So we are led to choose $\alpha_i, \alpha_{\bar i}$ in another way
(once again, the basic motivation for this choice is the presence of
at least one spin $1$ supercurrent with zero $u(1)$ charge after
splitting)
\begin{eqnarray}
\alpha_{1}=-1, \alpha_{\bar{1}}=1, \alpha_{2}=-2, \alpha_{\bar{2}}=0.
\label{eq:w32para}
\end{eqnarray}
It turns out that this is the right choice to produce the $N=2$ $W_3^{(2)}$
SCA precisely in the form given in \cite{IKS}, one of the surviving
supercurrents being the corresponding unconstrained
$N=2$ stress tensor. Actually, the choices \p{eq:w3para}, \p{eq:w32para}
are closely related to each other: the relevant $N=2$ stress tensors
differ by an improving term containing a spin $1/2$ fermionic
supercurrent. We will come back to this point later, while discussing
the secondary reduction of $N=2$ $W_3^{(2)}$ SCA.
Proceeding as before, we list in Table 4 the dimensions and
$u(1)$ charges of the constrained and gauge fixed supercurrents,
and in Table 5 indicate the supercurrents surviving the whole set of the
hamiltonian reduction second class constraints to be defined below (we
denote these latter supercurrents as ``surv. scs'').
\begin{eqnarray}
\begin{array}{cccccccccc}
\mbox{Table 4} \nonumber \\
\hline
u(1) & 0 & 0 & 1 & -1 & 1 & 0 & 0 & 0 & 0 \nonumber \\
\mbox{dim} & -1 & -1 & -\frac{1}{2} & -\frac{1}{2} & -\frac{1}{2}
& 0 & 0 & 0 & 0 \nonumber \\
\hline
\mbox{constr. scs} & {\cal J}_{13}^F & {\cal J}_{15}^B & {\cal J}_{12}^F &
{\cal J}_{43}^B & {\cal J}_{45}^F &
{\cal J}_{23}^F & {\cal J}_{14}^B & {\cal J}_{25}^B & {\cal J}_{42}^B \nonumber \\
\hline
\hline
\mbox{g.f. scs} & {\cal J}_{34}^B &{\cal J}_{52}^B & {\cal J}_{24}^B &
{\cal J}_{32}^F &
{\cal J}_{54}^F & {\cal J}_{35}^B & {\cal H}_{1}^F & {\cal H}_{2}^F & {\cal H}_{\bar{1}}^F \nonumber \\
\hline
\mbox{dim} & \frac{3}{2} & 1 & 1 & 1 & \frac{3}{2} & \frac{1}{2} &
\frac{1}{2} & \frac{1}{2} & \frac{1}{2} \nonumber \\
u(1) & 1 & 0 & 0 &
0 & -1 & 1 & 1 & 1 & -1 \nonumber \\
\hline
\end{array}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{cccccc||c}
\mbox{Table 5} \nonumber \\
\hline
\mbox{surv. scs} & {\cal J}_{53}^B & {\cal H}_{\bar{2}}^F & {\cal J}_{41}^B
& {\cal J}_{31}^F & {\cal J}_{51}^B &
{\cal J}_{21}^F \nonumber \\ \hline
\mbox{dim} & \frac{1}{2} & \frac{1}{2} & 1 & 2 & 2 & \frac{3}{2} \nonumber \\
u(1) & -1 & -1 & 0 & 0 & 0 & -1 \nonumber \\
\hline
\end{array}
\end{eqnarray}
In Table 5 and in similar Tables for other cases studied in this
Section we adopt the following convention: to the right from the
double vertical line we place those of the surviving
supercurrents (actually the single current ${\cal J}_{21}$ in the case at hand)
which are expressed through other ones by the remnants of the nonlinear
constraints (\ref{eq:cons}) after imposing the hamiltonian
reduction constraints.
These latter supercurrents themselves (they still can be constrained,
e.g., be chiral) are placed on the left.
From Table 4 we conclude that there are only three
bosonic affine supercurrents with both spin and $u(1)$ charge equal to zero,
namely, ${\cal J}_{14}$, ${\cal J}_{25}$ and ${\cal J}_{42}$. So we can put them
equal to $1$, while all the supercurrents with negative dimensions,
as in the previous examples, equal to zero. We also equate to zero
the fermionic supercurrent ${\cal J}_{23}$.
Thus the constraints we impose are of the form
\begin{eqnarray}
{\cal J}_{mn}^{constr}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & 0 \\
\ast & \ast & 0 & \ast & 1 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right).
\label{eq:consw32}
\end{eqnarray}
By plugging (\ref{eq:consw32}) into the nonlinear constraints
(\ref{eq:cons}), we can solve
one of them for ${\cal J}_{21}$ and express the latter in terms of ${\cal J}_{41}$
\begin{eqnarray}
{\cal J}_{21}=k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{1}} \right) {\cal J}_{41}.
\label{eq:j21}
\end{eqnarray}
As the next step we should fixe gauges. Gauge fixing procedure
can be performed using the same arguments as in the previous
examples and we eventually arrive at the following ${\cal J}_{mn}^{DS}$
\begin{eqnarray}
{\cal J}_{mn}^{DS}=
\left( \begin{array}{ccccc}
0 & 0 & 0 & 1 & 0 \\
k \bar{{\cal D}} {\cal J}_{41} & {\cal H}_{\bar{2}} & 0 & 0 & 1 \\
{\cal J}_{31} & 0 & 0 & 0 & 0 \\
{\cal J}_{41} & 1 & 0 & 0 & 0 \\
{\cal J}_{51} & 0 & {\cal J}_{53} & 0 & {\cal H}_{\bar{2}}
\end{array} \right).
\label{eq:dsw32}
\end{eqnarray}
Now we are ready to construct five independent gauge invariant
supercurrents by exploiting the general
procedure expounded in Section 3. After a lengthy but straightforward
computation we find
\begin{eqnarray}
\widetilde{{\cal J}_{53}} & = & {\cal J}_{53} - k \bar{{\cal D}} {\cal J}_{23} - k^3 \bar{{\cal D}}
{\cal J}_{13}' +
k {\cal H}_{\bar{2}} {\cal D} {\cal J}_{43} +
k^2 {\cal J}_{13} {\cal H}_{\bar{2}}' + k^2 {\cal J}_{15} {\cal J}_{53}' -
k \bar{{\cal D}} {\cal J}_{12} {\cal J}_{53} \nonumber \\
& & + k^2 \bar{{\cal D}} {\cal J}_{13} {\cal D} {\cal H}_{\bar{2}} -
k^2 \bar{{\cal D}} {\cal J}_{15} {\cal D} {\cal J}_{53} - k \bar{{\cal D}} {\cal J}_{45} {\cal J}_{53}
- k {\cal D} {\cal H}_{\bar{2}} {\cal J}_{43} + k^2 {\cal J}_{43}', \nonumber \\
\widetilde{{\cal H}_{\bar{2}}} & = & {\cal H}_{\bar{1}} + {\cal H}_{\bar{2}} +
k \bar{{\cal D}} {\cal J}_{14} + k \bar{{\cal D}} {\cal J}_{25} +
k^3 \bar{{\cal D}} {\cal J}_{15}' +
k^2 {\cal J}_{15} {\cal H}_{\bar{2}}' \nonumber \\
& & - k \bar{{\cal D}} {\cal J}_{12} {\cal H}_{\bar{2}} -
k^2 \bar{{\cal D}} {\cal J}_{15} {\cal D} {\cal H}_{\bar{2}} - k \bar{{\cal D}} {\cal J}_{45} {\cal H}_{\bar{2}}, \nonumber \\
\widetilde{{\cal J}_{41}} & = & {\cal J}_{24} + {\cal J}_{52} + k \bar{{\cal D}} {\cal H}_{1} + 2
k \bar{{\cal D}} {\cal H}_{2} +
k^3 \bar{{\cal D}} {\cal J}_{12}' -
k {\cal D} {\cal H}_{\bar{1}} + {\cal H}_{2} {\cal H}_{\bar{2}} \nonumber \\
& & - {\cal J}_{12} {\cal J}_{21} - {\cal J}_{13} {\cal J}_{31} +
{\cal J}_{14} {\cal J}_{41} + {\cal J}_{15} {\cal J}_{51} + {\cal J}_{35} {\cal J}_{53} +
k^2 {\cal J}_{14}' - k^2 {\cal J}_{42}', \nonumber \\
\widetilde{{\cal J}_{31}} & = & -2 {\cal J}_{31} + k \bar{{\cal D}} {\cal J}_{34} - k^3 \bar{{\cal D}}
{\cal J}_{35}' -
{\cal J}_{12} {\cal H}_{\bar{2}} {\cal J}_{31} + {\cal J}_{14} {\cal J}_{31} - k {\cal J}_{15}
{\cal D} {\cal H}_{\bar{2}} {\cal J}_{31} \nonumber \\
& & + {\cal J}_{25} {\cal J}_{31} + {\cal J}_{31} {\cal J}_{42} - {\cal J}_{32} {\cal J}_{41} - {\cal J}_{34}
{\cal H}_{\bar{2}} -
{\cal J}_{35} {\cal J}_{21} + {\cal J}_{35} {\cal H}_{\bar{2}} {\cal J}_{41} \nonumber \\
& & + k {\cal J}_{35} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}} +
k^2 {\cal J}_{35} {\cal H}_{\bar{2}}' - {\cal J}_{45} {\cal H}_{\bar{2}} {\cal J}_{31} -
k \bar{{\cal D}} {\cal J}_{35} {\cal J}_{41} - k^2 \bar{{\cal D}} {\cal J}_{35} {\cal D} {\cal H}_{\bar{2}} \nonumber \\
& & + k \bar{{\cal D}} {\cal J}_{45} {\cal J}_{31}
- k {\cal D} {\cal H}_{\bar{2}} {\cal J}_{32} +
k^2 {\cal J}_{15}' {\cal J}_{31} +
k^2 {\cal J}_{35}' {\cal H}_{\bar{2}} -
k^2 {\cal J}_{32}', \nonumber \\
\widetilde{{\cal J}_{51}} & = & -2 {\cal J}_{51} - k^3 \bar{{\cal D}} {\cal H}_{2}' - k {\cal D} {\cal J}_{21} +
k^3 {\cal D} {\cal H}_{\bar{1}}' - k {\cal H}_{\bar{1}} {\cal D} {\cal J}_{41} -
k {\cal H}_{\bar{2}} {\cal D} {\cal J}_{52} \nonumber \\
& & - 2 k {\cal H}_{\bar{2}} {\cal D} {\cal J}_{43} {\cal J}_{53}' -
2 k {\cal H}_{\bar{2}} {\cal D} {\cal J}_{43}' {\cal J}_{53} + {\cal H}_{2} {\cal J}_{21} -
{\cal H}_{2} {\cal H}_{\bar{2}} {\cal J}_{41} - k {\cal H}_{2} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}}
\nonumber \\
& & + k {\cal J}_{13} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{31} -
2 k^2 {\cal J}_{13} {\cal H}_{\bar{2}}' {\cal J}_{53}' -
2 k^2 {\cal J}_{13} {\cal H}_{\bar{2}}'' {\cal J}_{53} + {\cal J}_{14} {\cal J}_{51} +
k {\cal J}_{14} {\cal D} {\cal J}_{21} \nonumber \\
& & - k {\cal J}_{15} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{51}
+
k {\cal J}_{15} {\cal D} {\cal J}_{31} {\cal J}_{53} -
2 k^2 {\cal J}_{15} {\cal J}_{53}' {\cal J}_{53}' -
2 k^2 {\cal J}_{15} {\cal J}_{53}'' {\cal J}_{53}
\nonumber \\
& & - {\cal J}_{23} {\cal J}_{31} + {\cal J}_{25} {\cal J}_{51} + {\cal J}_{34} {\cal J}_{53} -
k {\cal J}_{35} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{53} - {\cal J}_{35} {\cal J}_{41} {\cal J}_{53} -
k {\cal J}_{35} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{53} \nonumber \\
& & - k^2 {\cal J}_{35} {\cal J}_{53}' -
{\cal J}_{41}' {\cal J}_{52} + {\cal J}_{42} {\cal J}_{51}
- k \bar{{\cal D}} {\cal H}_{2} {\cal J}_{41} -
k \bar{{\cal D}} {\cal J}_{12} {\cal J}_{51} - k^2 \bar{{\cal D}} {\cal J}_{12} {\cal D} {\cal J}_{21} \nonumber \\
& & + 4 k \bar{{\cal D}} {\cal J}_{12} {\cal J}_{53}' {\cal J}_{53} -
k^2 \bar{{\cal D}} {\cal J}_{13} {\cal D} {\cal J}_{31} -
2 k^2 \bar{{\cal D}} {\cal J}_{13} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{53}' -
2 k^2 \bar{{\cal D}} {\cal J}_{13} {\cal D} {\cal H}_{\bar{2}}' {\cal J}_{53} \nonumber \\
& & - k^2 \bar{{\cal D}} {\cal J}_{14} {\cal D} {\cal J}_{41} - k^2 \bar{{\cal D}} {\cal J}_{15} {\cal D} {\cal J}_{51} +
2 k^2 \bar{{\cal D}} {\cal J}_{15} {\cal D} {\cal J}_{53}' {\cal J}_{53} +
2 k^2 \bar{{\cal D}} {\cal J}_{15} {\cal J}_{53}' {\cal D} {\cal J}_{53} \nonumber \\
& & + 2 k \bar{{\cal D}} {\cal J}_{23} {\cal J}_{53}'
+
4 k \bar{{\cal D}} {\cal J}_{45} {\cal J}_{53}' {\cal J}_{53} +
2 k \bar{{\cal D}} {\cal J}_{12}' {\cal J}_{53} {\cal J}_{53} -
2 k^2 \bar{{\cal D}} {\cal J}_{13}' {\cal D} {\cal H}_{\bar{2}} {\cal J}_{53}
\nonumber \\
& & + 2 k^2 \bar{{\cal D}} {\cal J}_{15}' {\cal J}_{53} {\cal D} {\cal J}_{53} +
2 k \bar{{\cal D}} {\cal J}_{23}' {\cal J}_{53} +
2 k \bar{{\cal D}} {\cal J}_{45}' {\cal J}_{53} {\cal J}_{53} +
2 k^3 \bar{{\cal D}} {\cal J}_{13}'' {\cal J}_{53}
\nonumber \\
& & + 2 k {\cal D} {\cal H}_{\bar{2}} {\cal J}_{43} {\cal J}_{53}' +
2 k {\cal D} {\cal H}_{\bar{2}} {\cal J}_{43}' {\cal J}_{53} + k {\cal D} {\cal J}_{21} {\cal J}_{42} +
k {\cal D} {\cal J}_{31} {\cal J}_{43} + k {\cal D} {\cal J}_{32} {\cal J}_{53} \nonumber \\
& & + 2 k {\cal D} {\cal H}_{\bar{2}}' {\cal J}_{43} {\cal J}_{53} -
2 k {\cal H}_{\bar{2}}' {\cal D} {\cal J}_{43} {\cal J}_{53} -
2 k^2 {\cal H}_{2}' {\cal H}_{\bar{2}} +
k^2 {\cal J}_{12}' {\cal J}_{21} -
2 k^2 {\cal J}_{13}' {\cal H}_{\bar{2}}' {\cal J}_{53} \nonumber \\
& & - k^2 {\cal J}_{14}' {\cal J}_{41} -
2 k^2 {\cal J}_{15}' {\cal J}_{53}' {\cal J}_{53} -
2 k^2 {\cal J}_{35}' {\cal J}_{53} -
2 k^2 {\cal J}_{43}' {\cal J}_{53}' -
2 {\cal J}_{53}' {\cal J}_{53} \nonumber \\
& & -k^2 {\cal H}_{2} {\cal H}_{\bar{2}}'
-k {\cal J}_{21} {\cal D} {\cal J}_{42}
+ 2 k^3 \bar{{\cal D}} {\cal J}_{13}' {\cal J}_{53}'
+ k {\cal D} {\cal H}_{\bar{1}} {\cal J}_{41}
- 2 k^2 {\cal J}_{43}'' {\cal J}_{53} \nonumber \\
& & - k^2 {\cal J}_{24}' -
k^2 {\cal J}_{52}' - k^4 {\cal J}_{14}''
- \frac{k^2}{2} [ {\cal D}, \bar{{\cal D}} ]
\widetilde{{\cal J}_{41}} + 2 \widetilde{{\cal J}_{53}'} \widetilde{{\cal J}_{53}} +
\frac{k^2}{2} \widetilde{{\cal J}_{41}}'.
\label{eq:tilw32}
\end{eqnarray}
It is easy to verify that these gauge invariant supercurrents
satisfy the condition (\ref{eq:tilj}).
Note that in the last three terms of $\widetilde{{\cal J}_{51}}$, in order
to shorten this expression, we kept
$\widetilde{{\cal J}_{41}}, \widetilde{{\cal J}_{53}}$ in their implicit form.
Now it is direct to calculate the star SOPEs between them using the rule
(3.9) and the explicit expressions (\ref{eq:tilw32}) and (\ref{eq:qq}).
The $N=2$ stress tensor is given by
\begin{eqnarray}
{\cal T}= -\frac{1}{k} \widetilde{{\cal J}_{41}} = -\frac{1}{k} {\cal J}_{41},
\label{eq:Tw32}
\end{eqnarray}
where the second equality is fulfilled on the shell of constraints.
It has the central charge $-2 k$ and coincides
with ${\cal T}_{sug}$ (\ref{eq:sugsl32}), (\ref{eq:w32para})
on the constraints shell.
After the redefinitions
\begin{eqnarray}
{\cal J}_{53} \rightarrow {\cal J}_{53}, \;\;
{\cal H}_{\bar{2}} \rightarrow {\cal H}_{\bar{2}}, \;\;
{\cal J}_{31} \rightarrow \frac{1}{k^3} {\cal J}_{31}, \;\;
{\cal J}_{51} \rightarrow \frac{1}{k^3} {\cal J}_{51}\;,
\end{eqnarray}
all the supercurrents except for ${\cal H}_{\bar{2}}$ are superprimary with
respect to the stress tensor (\ref{eq:Tw32}) (see eq. (\ref{eq:JJ})),
and have the spins $1/2, 1/2, 2$ and $2$, respectively, while
${\cal H}_{\bar{2}}$ is quasi superprimary
\begin{eqnarray}
{\cal T} (Z_{1}) {{\cal H}}_{\bar{2}} (Z_{2})=
\frac{\theta_{12}}{z_{12}^2} 2 k +
\left[ \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \frac{1}{2} -
\frac{\theta_{12}}{z_{12}} {\cal D} + \frac{\bar{\theta}_{12}}{z_{12}} \bar{{\cal D}} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \partial -
\frac{1}{z_{12}} \right] {{\cal H}}_{\bar{2}}.
\label{eq:Th2}
\end{eqnarray}
Finally, the remaining set of SOPEs is as follows (from now on, we omit the
index ``*'', keeping in mind that all such SOPEs are
computed according to the rule (3.9))
\begin{eqnarray}
{\cal H}_{\bar{2}}(Z_{1}) {\cal J}_{53} (Z_{2}) & = & \frac{\theta_{12}}{z_{12}}
{\cal J}_{53}, \nonumber \\
{\cal H}_{\bar{2}}(Z_{1}) {\cal J}_{31} (Z_{2}) & = & -\frac{\theta_{12}}{z_{12}}
{\cal J}_{31}, \nonumber \\
{\cal H}_{\bar{2}}(Z_{1}) {\cal J}_{51} (Z_{2}) & = &
-\frac{\theta_{12}}{z_{12}^3} 2 k + \frac{\theta_{12} {\bar{\theta}}_
{12}}{
z_{12}^3} {\cal H}_{\bar{2}}+
\frac{1}{z_{12}^2} {\cal H}_{\bar{2}} + \frac{\theta_{12}}{z_{12}^2}
\left[ {\cal T} -
{\cal D} {\cal H}_{\bar{2}} \right] \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[ \frac{1}{2} \bar{{\cal D}} {\cal T} +
\frac{1}{2 k} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}} -
\frac{1}{2 k} {\cal T} {\cal H}_{\bar{2}} +
\frac{1}{2} {\cal H}_{\bar{2}}'\right] \nonumber \\
& & +
\frac{1}{z_{12}} \left[ \bar{{\cal D}} {\cal T} + \frac{1}{k} {\cal H}_{\bar{2}}
{\cal D} {\cal H}_{\bar{2}} - \frac{1}{k}
{\cal T} {\cal H}_{\bar{2}} +
{\cal H}_{\bar{2}}' \right], \nonumber \\
{\cal J}_{53}(Z_{1}) {\cal J}_{31}(Z_{2}) & = &
\frac{\theta_{12}}{z_{12}^3} 2 k - \frac{\theta_{12} {\bar{\theta}}_
{12}}{
z_{12}^3} {\cal H}_{\bar{2}}
-\frac{1}{z_{12}^2} {\cal H}_{\bar{2}}
+ \frac{\theta_{12}}{z_{12}^2} \left[-
{\cal T} +
{\cal D} {\cal H}_{\bar{2}} \right] \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[ -
\frac{1}{2} \bar{{\cal D}} {\cal T} -
\frac{1}{2 k} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}} +
\frac{1}{2 k} {\cal T} {\cal H}_{\bar{2}} -
\frac{1}{2} {\cal H}_{\bar{2}}' \right] \nonumber \\
& & + \frac{1}{z_{12}} \left[
- \bar{{\cal D}} {\cal T} - \frac{1}{k} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}} +
\frac{1}{k} {\cal T} {\cal H}_{\bar{2}}-
{\cal H}_{\bar{2}}'\right]-\frac{\theta_{12}}{z_{12}} {\cal J}_{51}, \nonumber \\
{\cal J}_{53}(Z_{1}) {\cal J}_{51}(Z_{2}) & = &
\frac{\theta_{12} {\bar{\theta}}_{12}}{z_{12}^3} {\cal J}_{53}+
\frac{1}{z_{12}^2} {\cal J}_{53} - \frac{\theta_{12}}{z_{12}^2}
{\cal D} {\cal J}_{53} \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[
\frac{1}{2 k} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{53} -
\frac{1}{2 k} {\cal T} {\cal J}_{53} +
\frac{1}{2 k} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{53} +
\frac{1}{2} {\cal J}_{53}' \right] \nonumber \\
& & + \frac{1}{z_{12}} \left[
\frac{1}{k} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{53} - \frac{1}{k}
{\cal T} {\cal J}_{53} +
\frac{1}{k} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{53} +
{\cal J}_{53}' \right], \nonumber \\
{\cal J}_{31}(Z_{1}) {\cal J}_{51}(Z_{2}) & = &
-\frac{1}{z_{12}^2} 2 {\cal J}_{31} - \frac{\bar{\theta}_{12}}{z_{12}^2} \frac{1}{k} {\cal H}_{\bar{2}}
{\cal J}_{31} -
\frac{{\theta}_{12}}{z_{12}^2} {\cal D} {\cal J}_{31} \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[- \frac{1}{k} {\cal T} {\cal J}_{31} +
\frac{3}{k} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{31} +
\frac{3}{2} {\cal J}_{31}' \right]
- \frac{1}{z_{12}} {\cal J}_{31}' \nonumber \\
& & +
\frac{\bar{\theta}_{12}}{z_{12}} \left[ \frac{1}{k^2} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{31} -
\frac{1}{k^2} {\cal T} {\cal H}_{\bar{2}} {\cal J}_{31} +
\frac{1}{k} \bar{{\cal D}} {\cal T} {\cal J}_{31} -
\frac{1}{k} {\cal H}_{\bar{2}}' {\cal J}_{31} \right] \nonumber \\
& & + \frac{\theta_{12}}{z_{12}} \left[-
{\cal D} {\cal J}_{31}' + \frac{1}{k} {\cal T} {\cal D} {\cal J}_{31} -
\frac{1}{k} {\cal D} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{31} - \frac{1}{k}
{\cal D} {\cal T} {\cal J}_{31} \right] \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[ -\frac{1}{k} {\cal T} {\cal J}_{31}' +
\frac{1}{k} \bar{{\cal D}} {\cal T} {\cal D} {\cal J}_{31} +
\frac{2}{k} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{31}' +
\frac{1}{k^2} {\cal D} {\cal T} {\cal H}_{\bar{2}} {\cal J}_{31} \right. \nonumber \\
& & \left.
+ \frac{2}{k} {\cal D} {\cal H}_{\bar{2}}' {\cal J}_{31} +
{\cal J}_{31}'' \right], \nonumber \\
{\cal J}_{51} (Z_{1}) {\cal J}_{51} (Z_{2}) & = &
-\frac{1}{z_{12}^2} 2 {\cal J}_{51} - \frac{\bar{\theta}_{12}}{z_{12}^2} \left[ \frac{1}{k}
{\cal H}_{\bar{2}} {\cal J}_{51} +
\frac{1}{k} {\cal J}_{31} {\cal J}_{53} \right] - \frac{{\theta}_{12}}{z_{12}^2} {\cal D}
{\cal J}_{51}
- \frac{1}{z_{12}} {\cal J}_{51}' \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[ -
\frac{3}{k} {\cal J}_{31} {\cal D} {\cal J}_{53} -
\frac{1}{k} {\cal T} {\cal J}_{51} +
\frac{3}{k} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{51} +
\frac{3}{2} {\cal J}_{51}' \right]
\nonumber \\
& & +
\frac{\bar{\theta}_{12}}{z_{12}} \left[ -\frac{1}{k^2} {\cal H}_{\bar{2}} {\cal J}_{31} {\cal D}
{\cal J}_{53} +
\frac{1}{k^2} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}}
{\cal J}_{51} -
\frac{1}{k} {\cal J}_{31} {\cal J}_{53}'
\right. \nonumber \\
& & - \frac{1}{k^2} {\cal T} {\cal J}_{31} {\cal J}_{53} +
\frac{1}{k} \bar{{\cal D}} {\cal T} {\cal J}_{51} +
\frac{1}{k^2} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{31} {\cal J}_{53} -
\frac{1}{k} {\cal H}_{\bar{2}}' {\cal J}_{51} \nonumber \\
& & \left. - \frac{1}{k^2} {\cal T} {\cal H}_{\bar{2}} {\cal J}_{51} \right] \nonumber \\
& & + \frac{\theta_{12}}{z_{12}} \left[-
{\cal D} {\cal J}_{51}' + \frac{1}{k} {\cal T} {\cal D} {\cal J}_{51} -
\frac{1}{k} {\cal D} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{51}
+
\frac{1}{k} {\cal D} {\cal J}_{31} {\cal D} {\cal J}_{53} \right. \nonumber \\
& & \left. - \frac{1}{k} {\cal D} {\cal T} {\cal J}_{51}
\right] \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[-
\frac{2}{k} {\cal J}_{31} {\cal D} {\cal J}_{53}' -
\frac{1}{k} {\cal T} {\cal J}_{51}' +
\frac{1}{k} \bar{{\cal D}} {\cal T} {\cal D} {\cal J}_{51} +
\frac{2}{k} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{51}'
\right. \nonumber \\
& &
+ \frac{1}{k^2} {\cal D} {\cal T} {\cal H}_{\bar{2}} {\cal J}_{51} +
\frac{1}{k^2} {\cal D} {\cal T} {\cal J}_{31} {\cal J}_{53} +
\frac{2}{k} {\cal D} {\cal H}_{\bar{2}}' {\cal J}_{51} +{\cal J}_{51}'' \nonumber \\
& & \left. -
\frac{2}{k} {\cal J}_{31}' {\cal D} {\cal J}_{53} \right].
\end{eqnarray}
So we end up with the following five $N=2$ supercurrents:
a general spin $1$
${\cal T}$, spin $1/2$ antichiral fermionic ${\cal H}_{\bar{2}}$ and bosonic
${\cal J}_{53}$, constrained spin $2$ fermionic ${\cal J}_{31}$ and
bosonic ${\cal J}_{51}$ ones.
We also write down the constraints which stem from the original
nonlinear constraints on the affine supercurrents
\begin{eqnarray}
\bar{{\cal D}} {\cal H}_{\bar{2}} = 0, & \;\; & \bar{{\cal D}} {\cal J}_{53} = 0, \nonumber \\
\left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{2}} \right)
{\cal J}_{31} = 0, & \;\; & \left( \bar{{\cal D}} -
\frac{1}{k} {\cal H}_{\bar{2}} \right) {\cal J}_{51} -
\frac{1}{k} {\cal J}_{31} {\cal J}_{53} = 0.
\label{eq:w32con}
\end{eqnarray}
By construction, all the above SOPEs are compatible with these constraints.
These SOPEs and constraints constitute the superfield description of $N=2$
$W_3^{(2)}$ superalgebra given in \cite{IKS}.
As shown in \cite{IKS}, we can obtain $N=2$ $W_{3}$ SCA from $N=2$
$W_{3}^{(2)}$ SCA by means of secondary hamiltonian
reduction \cite{DFRS} (by the primary hamiltonian reduction we
mean the one which proceeds directly from affine (super)algebra).
With respect to the new stress tensor ${\cal T}_{\mbox{new}}$,
\begin{eqnarray}
{\cal T}_{\mbox{new}}= {\cal T}+ {\cal D} {\cal H}_{\bar{2}}
\label{eq:w32newT}
\end{eqnarray}
${\cal J}_{53}$ has vanishing spin and $u(1)$ charge.
Then we can put nonzero constraint on ${\cal J}_{53}$
\begin{eqnarray}
{\cal J}_{53}-1 =0
\end{eqnarray}
and gauge away ${\cal H}_{\bar{2}}$
\begin{eqnarray}
{\cal H}_{\bar{2}}=0.
\end{eqnarray}
These additional constraints are consistent with the first and
second of eqs. (\ref{eq:w32con}), respectively.
One observes that now ${\cal J}_{31}$ is expressed from the fourth of
eqs. (\ref{eq:w32con}) as
\begin{eqnarray}
{\cal J}_{31}= k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{
\bar{2}} \right) {\cal J}_{51}\;,
\end{eqnarray}
after which the third of eqs. (\ref{eq:w32con}) is satisfied identically.
Then we are left with the same ${\cal J}_{mn}^{DS}$ as in (\ref{eq:w3ds}).
Thus the surviving independent supercurrents are ${\cal T}_{\mbox{new}}$ and $
{\cal J}_{51}$ and it remains to construct the appropriate gauge invariant
supercurrents and to compute their SOPEs using the rule (3.9).
The dimension of ${\cal J}_{51}$ and its $u(1)$ charge
with respect to ${\cal T}_{\mbox{new}}$ are the same as in Table 5,
i.e. 2 and 0, which are characteristic of the second supercurrent of
$N=2$ $W_3$ SCA.
According to \cite{IKS}, the resulting superalgebra is precisely
the $N=2$ $W_{3}$ SCA \cite{LPRSW}. Of course, we could arrive
at the same SOPEs directly in the framework of the primary hamiltonian
reduction procedure described in the previous Subsection.
Let us remark that ${\cal T}_{\mbox{new}}$ \p{eq:w32newT}
exactly corresponds to the previous choice of the
splitting parameters (\ref{eq:w3para}), in the sense that the dimensions
and $u(1)$ charges of all the supercurrents with respect to it
are the same as in Subsect. 5.1.
This implies that the bases (\ref{eq:w32para})
and (\ref{eq:w3para}) are related through the shift
$\sim {\cal D} {\cal H}_{\bar{2}}$ of the respective $N=2$ stress tensors.
We could equally derive $N=2$ $W_3^{(2)}$ SCA sticking
to the choice (\ref{eq:w3para}) and relaxing one of the constraints
of units (on ${\cal J}_{53}$) directly in the supermatrix (\ref{eq:w3cons}).
However, in the corresponding
basis the $N=2$ $W_3^{(2)}$ supercurrents are even not quasi superprimary.
To put this superalgebra in the standard form given in
\cite{IKS} one should pass to the stress tensor ${\cal T}$ (\ref{eq:Tw32})
by the relation \p{eq:w32newT}.
It is worth to notice that one can produce eight reduction constraints by
relaxing of the constraint on ${\cal J}_{23}$ in the supermatrix (5.7). Then the
surviving supercurrents have extra two supercurrents ${\cal J}_{23}, {\cal J}_{35}$
in addition to the superfield contents of $N=2$ $W_{3}^{(2)}$ SCA.
In next Subsection we will consider more examples of extended $N=2$
conformal superalgebras obtained from $N=2$ $sl(3|2)^{(1)}$ via
$N=2$ superfield hamiltonian reduction.
\subsection{\bf New Extended $N=2$ SCAs}
From now on we will concentrate on those examples of
hamiltonian reduction in $N=2$ superspace which generate
new extended $N=2$ SCAs.
Next natural step is to consider the cases in which the number of
the reduction constraints is less than nine. Let us first
describe the case with five constraints (this number is the minimal one
at which the constraints can still be chosen to be first-class).
As before, the reason why we choose the specific values for
splitting parameters as below stems from the demand that among
the surviving supercurrents there is at least one bosonic supercurrent
with spin $1$ and $u(1)$ charge zero.
For the choice
\begin{eqnarray}
\alpha_{1}=-1, \;\; \alpha_{\bar{1}}=0, \;\; \alpha_{2}=0,
\;\; \alpha_{\bar{2}}=1
\end{eqnarray}
we list the dimensions and $u(1)$ charges of supercurrents in
Tables 6 and 7.
\begin{eqnarray}
\begin{array}{cccccc}
\mbox{Table 6} \nonumber \\
\hline
u(1) & -1 & 0 & 2 & 0 & 0 \nonumber \\
\mbox{dim} & -\frac{1}{2} & 0 & 0 & 0 & 0 \nonumber \\
\hline
\mbox{constr. scs} & {\cal J}_{54}^F & {\cal J}_{14}^B & {\cal J}_{24}^B &
{\cal J}_{34}^B &
{\cal J}_{53}^B \nonumber \\
\hline
\hline
\mbox{g.f. scs} & {\cal J}_{43}^B & {\cal H}_{1}^F &{\cal J}_{12}^F &
{\cal J}_{13}^F &
{\cal H}_{\bar{2}}^F \nonumber \\ \hline
\mbox{dim} & 1 & \frac{1}{2} & \frac{1}{2} & \frac{1}{2} & \frac{1}{
2} \nonumber \\
u(1) & 0 & 1 & -1 & 1 &-1 \nonumber \\
\hline
\end{array}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{cccccccccccc||ccc}
\mbox{Table 7} \nonumber \\
\hline
\mbox{surv. scs}&{\cal J}_{51}^B&{\cal J}_{52}^B&{\cal J}_{21}^F&{\cal J}_{23}^F&
{\cal H}_{\bar{1}}^F&{\cal H}_{2}^F&{\cal J}_{15}^B&{\cal J}_{35}^B&{\cal J}_{41}^B&{\cal J}_{42}^B&
{\cal J}_{25}^B&{\cal J}_{31}^F&{\cal J}_{32}^F&{\cal J}_{45}^F\nonumber \\ \hline
\mbox{dim} & 0 & 0 & \frac{1}{2} & \frac{1}{2} & \frac{1}{2}
& \frac{1}{2} & 1 & 1 & 1 & 1 & 1
& \frac{1}{2} & \frac{1}{2} & \frac{3}{2} \nonumber \\
u(1) & 0 & -2 & 1 & -1 & -1 & 1 & 0 & 0 & 0 & -2 & 2 & -1
& -3 & 1 \nonumber \\
\hline
\end{array}
\end{eqnarray}
We can choose the appropriate constraints as follows
\begin{eqnarray}
{\cal J}_{mn}^{constr}=
\left( \begin{array}{ccccc}
\ast & \ast & \ast & 1 & \ast \\
\ast & \ast & \ast & 0 & \ast \\
\ast & \ast & \ast & 0 & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & 1 & 0 & \ast
\end{array} \right).
\label{eq:conso5}
\end{eqnarray}
One observes that it is not a subset of constraints discussed in
previous Subsections, (\ref{eq:w3cons}), (\ref{eq:consw32}).
Further we fix the gauge according to Table 6 and
quote the surviving supercurrents in Table 7.
There are three supercurrents expressible at expense of the
remaining ones, which can be seen by substituting \p{eq:conso5}
into the nonlinear constraints
(\ref{eq:cons})
\begin{eqnarray}
{\cal J}_{31} & = & k \left( \bar{{\cal D}} -\frac{1}{k} \left( {\cal H}_{\bar{1}} +{\cal H}_{\bar{2}}
\right) \right) {\cal J}_{51} - {\cal J}_{21} {\cal J}_{52}, \nonumber \\
{\cal J}_{32} & = & k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{2}} \right) {\cal J}_{52},
\nonumber \\
{\cal J}_{45} & = & k \left( {\cal D} +\frac{1}{k} {\cal H}_{1} \right) {\cal J}_{15} -
{\cal J}_{12} {\cal J}_{25}-{\cal J}_{13} {\cal J}_{35}.
\end{eqnarray}
Thus we come to the following ${\cal J}_{mn}^{DS}$
\begin{eqnarray}
{\cal J}_{mn}^{DS}=
\left( \begin{array}{ccccc}
{\cal H}_{\bar{1}} & 0 & 0 & 1 & {\cal J}_{15} \\
{\cal J}_{21} & 0 & {\cal J}_{23} & 0 & {\cal J}_{25} \\
k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{1}} \right) {\cal J}_{51} -
{\cal J}_{21} {\cal J}_{52} & k \bar{{\cal D}} {\cal J}_{52} & {\cal H}_{2} & 0 & {\cal J}_{35} \\
{\cal J}_{41} & {\cal J}_{42} & 0 & {\cal H}_{\bar{1}} & k {\cal D} {\cal J}_{15} \\
{\cal J}_{51} & {\cal J}_{52} & 1 & 0 & {\cal H}_{2} \\
\end{array} \right).
\label{eq:DS5}
\end{eqnarray}
The supercurrents ${\cal J}_{51}, {\cal J}_{52},$ and ${\cal J}_{15}$ remain unconstrained.
Once we know the gauge invariant supercurrents,
it is straightforward to deduce their algebra.
The consruction of these gauge invariant quantities is the crucial
(and most difficult) step of our approach.
With the above choice of five constraints, it is rather lengthy and
cumbersome to find correct ansatz for gauge invariant supercurrents
because two spin $0$ supercurrents ${\cal J}_{51}, {\cal J}_{52}$ are present
(let us remind that $\widetilde{{\cal J}_{\alpha \beta}}$ are some
nonlinear functionals of ${\cal J}_{\alpha \beta}$ and their derivatives).
As the first step we write down $\widetilde{{\cal J}_{\alpha
\beta}}$ as a lowest order monomial in ${\cal J}_{51}, {\cal J}_{52}$,
and check whether it
satisfies the conditions
(\ref{eq:tilj}), (\ref{eq:req}). If this is not the case, we include
next order terms in ${\cal J}_{51}, {\cal J}_{52}$, etc., until the
conditions (\ref{eq:tilj}), (\ref{eq:req}) are satisfied.
Finally we obtain the gauge invariant
supercurrents $\widetilde{
{\cal J}_{ \alpha \beta}}$ in the following form
\begin{eqnarray}
\widetilde{{\cal J}_{51}} & = &
{\cal J}_{34} + 2 {\cal J}_{51} - k {\cal D} {\cal J}_{54} + {\cal H}_{2} {\cal J}_{54}+
{\cal J}_{23} {\cal J}_{52} {\cal J}_{54} +{\cal J}_{24} {\cal J}_{52} - {\cal J}_{51} {\cal J}_{53}, \nonumber \\
\widetilde{{\cal J}_{52}} & = &
2 {\cal J}_{52} - {\cal J}_{14} {\cal J}_{52}, \nonumber \\
\widetilde{{\cal J}_{21}} & = &
{\cal J}_{21} + k \bar{{\cal D}} {\cal J}_{24} + {\cal J}_{14} {\cal J}_{21} -
{\cal J}_{21} {\cal J}_{53} + {\cal J}_{23} {\cal H}_{\bar{1}} {\cal J}_{54} -
{\cal J}_{24} {\cal H}_{\bar{1}} + k \bar{{\cal D}} {\cal J}_{23} {\cal J}_{54}, \nonumber \\
\widetilde{{\cal J}_{23}} & = &
{\cal J}_{23} + {\cal J}_{23} {\cal J}_{14} - {\cal J}_{23} {\cal J}_{53},
\nonumber \\
\widetilde{{\cal H}_{\bar{1}}} & = &
{\cal H}_{\bar{1}} + k \bar{{\cal D}} {\cal J}_{14}, \nonumber \\
\widetilde{{\cal H}_{2}} & = &
{\cal H}_{2} + k {\cal D} {\cal J}_{53}, \nonumber \\
\widetilde{{\cal J}_{15}} & = &
{\cal J}_{43} - k \bar{{\cal D}} {\cal J}_{13} + {\cal J}_{15} {\cal J}_{53}, \nonumber \\
\widetilde{{\cal J}_{35}} & = &
-k {\cal D} {\cal H}_{\bar{2}} + {\cal H}_{2} {\cal H}_{\bar{2}} - k {\cal H}_{2}
\bar{{\cal D}} {\cal J}_{14} -
{\cal H}_{2} {\cal J}_{15} {\cal J}_{54} \nonumber \\
& & - {\cal J}_{15} {\cal J}_{34} + k {\cal J}_{15} {\cal D} {\cal J}_{54} +
{\cal J}_{35} {\cal J}_{53} + {\cal J}_{45} {\cal J}_{54} - k^2 {\cal J}_{14}', \nonumber \\
\widetilde{{\cal J}_{41}} & = & k \bar{{\cal D}} {\cal H}_{1} + k {\cal H}_{\bar{1}} {\cal D} {\cal J}_{53} +
{\cal H}_{1} {\cal H}_{\bar{1}} - {\cal J}_{12} {\cal J}_{21} -
{\cal J}_{13} {\cal H}_{\bar{2}} {\cal J}_{51} - {\cal J}_{13} {\cal J}_{31} {\cal J}_{53}
\nonumber \\
& & + {\cal J}_{14} {\cal J}_{41} + {\cal J}_{23} {\cal J}_{42} {\cal J}_{
54} +
{\cal J}_{24} {\cal J}_{42} - {\cal J}_{43} {\cal J}_{51} + k \bar{{\cal D}} {\cal J}_{13} {\cal J}_{51} +
k^2 {\cal J}_{53}', \nonumber \\
& & - {\cal J}_{13} {\cal J}_{41} {\cal J}_{54} \nonumber \\
\widetilde{{\cal J}_{42}} & = &
-k \bar{{\cal D}} {\cal J}_{12} - {\cal J}_{13} {\cal H}_{\bar{2}} {\cal J}_{52} -
{\cal J}_{13} {\cal J}_{32} {\cal J}_{53} -
{\cal J}_{13} {\cal J}_{42} {\cal J}_{54} + {\cal J}_{42} {\cal J}_{53} - {\cal J}_{43} {\cal J}_{52} \nonumber \\
& & +
k \bar{{\cal D}} {\cal J}_{13} {\cal J}_{52}, \nonumber \\
\widetilde{{\cal J}_{25}} & = &
{\cal J}_{14} {\cal J}_{25} - {\cal J}_{15} {\cal J}_{24} + {\cal J}_{23}
{\cal H}_{\bar{2}} - k {\cal J}_{23} \bar{{\cal D}} {\cal J}_{14} +
k {\cal J}_{23} \bar{{\cal D}} {\cal J}_{53} - {\cal J}_{23} {\cal J}_{15} {\cal J}_{54}.
\end{eqnarray}
We also construct $N=2$ stress tensor
\begin{eqnarray}
{\cal T} =-\frac{1}{k} \left[ {\cal J}_{35} +{\cal J}_{41} +
{\cal H}_{2} {\cal H}_{\bar{1}} +
{\cal J}_{15} {\cal J}_{51} -k
{\cal J}_{23} \bar{{\cal D}} {\cal J}_{52} + {\cal J}_{25}
{\cal J}_{52} \right]
\label{eq:T}
\end{eqnarray}
with central charge $2k$. The remnants of nonlinear irreducibility
constraints are given by
\begin{eqnarray}
\bar{{\cal D}} {\cal H}_{\bar{1}} = 0, & \;\; & {\cal D} {\cal H}_{2} = 0, \nonumber \\
\left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{1}} \right)
{\cal J}_{21} = 0, & \;\; & \left( {\cal D} +\frac{1}{k} {\cal H}_{2} \right)
{\cal J}_{23} = 0, \nonumber \\
\left( {\cal D} - \frac{1}{k} {\cal H}_{2} \right)
{\cal J}_{35} = 0, & \;\; & \left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{1}} \right)
{\cal J}_{41}+
\frac{1}{k} {\cal J}_{21} {\cal J}_{42} = 0, \nonumber \\
\bar{{\cal D}} {\cal J}_{42} = 0, & \;\; & {\cal D} {\cal J}_{25} -\frac{1}{k} {\cal J}_{23}
{\cal J}_{35} = 0.
\end{eqnarray}
The above gauge invariant supercurrents form some extended
$N=2$ SCA, in particular, the stress tensor \p{eq:T} generates
the standard $N=2$ SCA. Here we do not present this superalgebra
explicitly and leave its study to the future. The reason is that
it does not meet one of the criterions by which we limited from the
beginning our study in this paper. Namely, the
$N=2$ stress tensor (\ref{eq:T}) is constrained
because the linear terms in (\ref{eq:T}) ${\cal J}_{35}, {\cal J}_{41}$
are constrained. The only unconstrained bosonic supercurrent with
the spin and $u(1)$ charge appropriate for $N=2$ stress tensor,
${\cal J}_{15}$, enters nonlinearly into ${\cal T}$ (\ref{eq:T}),
so eq. (\ref{eq:T}) does not imply an invertible relation between
${\cal T}$ and ${\cal J}_{15}$.
We would like to mention that the supercurrents
\begin{eqnarray}
{\cal H}_{2}, \;\;\
{\cal H}_{\bar{1}}, \;\;\
{\cal J}_{35}-k \bar{{\cal D}} {\cal H}_{2}
\end{eqnarray}
are quasi superprimary, and
\begin{eqnarray}
{\cal J}_{51}, \;\;
{\cal J}_{52}, \;\;
{\cal J}_{21}, \;\;
{\cal J}_{23}, \;\;
{\cal J}_{15}, \;\;
{\cal J}_{42}, \;\;
{\cal J}_{25}-\frac{k}{2} \bar{{\cal D}} {\cal J}_{23}
\end{eqnarray}
are superprimary with respect to ${\cal T}$ (\ref{eq:T}).
Now we turn to another choice of five constraints
which leads to an unconstrained $N=2$ stress tensor and so seems to be
more interesting. For
\begin{eqnarray}
\alpha_{1}=-1, \alpha_{\bar{1}}=1, \alpha_{2}=0, \alpha_{\bar{2}}=0
\label{eq:new5para}
\end{eqnarray}
we have the spins and $u(1)$ charges as is given in Tables 8, 9.
\begin{eqnarray}
\begin{array}{cccccc}
\mbox{Table 8} \nonumber \\
\hline
u(1) & 1 & 2 & 0 & 0 & 1 \nonumber \\
\mbox{dim} & -\frac{1}{2} & 0 & 0 & 0 & \frac{1}{2} \nonumber \\
\hline
\mbox{constr. scs} & {\cal J}_{12}^F & {\cal J}_{13}^F & {\cal J}_{42}^B &
{\cal J}_{14}^B & {\cal J}_{43}^B \nonumber \\
\hline
\hline
\mbox{g.f. scs} & {\cal J}_{24}^B &{\cal J}_{34}^B & {\cal H}_{\bar{1}}^F
& {\cal H}_{1}^F &
{\cal J}_{32}^F \nonumber \\ \hline
\mbox{dim} & 1 & \frac{1}{2} & \frac{1}{2} & \frac{1}{2} & 0 \nonumber \\
u(1) & 0 & -1 & -1 & 1 & -2 \nonumber \\
\hline
\end{array}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{cccccccccccc||ccc}
\mbox{Table 9} \nonumber \\
\hline
\mbox{surv. scs} & {\cal J}_{15}^B & {\cal J}_{52}^B & {\cal J}_{53}^B
& {\cal J}_{35}^B &
{\cal H}_{2}^F & {\cal H}_{\bar{2}}^F & {\cal J}_{25}^B & {\cal J}_{51}^B & {\cal J}_{23}^F &
{\cal J}_{41}^B &
{\cal J}_{31}^F & {\cal J}_{54}^F & {\cal J}_{45}^F & {\cal J}_{21}^F \nonumber \\ \hline
\mbox{dim} & 0 & 0 & \frac{1}{2} & \frac{1}{2} & \frac{1}{2} & \frac{1}{2}
& 1 & 1 &
1 & 1 & 1 &
\frac{1}{2} & \frac{1}{2} & \frac{3}{2} \nonumber \\
u(1) & 2 & -2 & -1 & 1 & 1 & -1 & 2 & -2 & 2 & 0 & -2 & -3
& 3 & -1 \nonumber \\
\hline
\end{array}
\end{eqnarray}
With this choice of parameters, we impose the following constraints:
\begin{eqnarray}
{\cal J}_{mn}^{constr}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right).
\label{eq:consn5}
\end{eqnarray}
Then, fixing gauges according to Table 8, for surviving supercurrents
we have Table 9.
In Table 9, last three supercurrents are expressed through
the remaining ones by the relations
\vspace{2cm}
\begin{eqnarray}
{\cal J}_{45} & = & k \left( {\cal D} +\frac{1}{k} {\cal H}_{1} \right) {\cal J}_{15}, \nonumber \\
{\cal J}_{21} & = & k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{1}} \right) {\cal J}_{41},
\nonumber \\
{\cal J}_{54} & = & k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{2}} \right) {\cal J}_{52} -
{\cal J}_{32} {\cal J}_{53}.
\end{eqnarray}
The supercurrents ${\cal J}_{15}, {\cal J}_{41},$ and ${\cal J}_{52}$ are unconstrained.
Then one gets the following ${\cal J}_{mn}^{DS}$
\begin{eqnarray}
{\cal J}_{mn}^{DS}=
\left( \begin{array}{ccccc}
0 & 0 & 0 & 1 & {\cal J}_{15} \\
k \bar{{\cal D}} {\cal J}_{41} & {\cal H}_{\bar{2}} & {\cal J}_{23} & 0 & {\cal J}_{25} \\
{\cal J}_{31} & 0 & {\cal H}_{2} & 0 & {\cal J}_{35} \\
{\cal J}_{41} & 1 & 0 & 0 & k {\cal D} {\cal J}_{15} \\
{\cal J}_{51} & {\cal J}_{52} & {\cal J}_{53} & k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{2}}
\right)
{\cal J}_{52} & {\cal H}_{2}+{\cal H}_{\bar{2}}
\end{array} \right).
\label{eq:dsds}
\end{eqnarray}
The gauge invariant supercurrents are given by the following
expressions
\begin{eqnarray}
\widetilde{{\cal J}_{15}} & = &
2 {\cal J}_{15} - {\cal J}_{15} {\cal J}_{42} + k \bar{{\cal D}} {\cal J}_{12} {\cal J}_{15}, \nonumber \\
\widetilde{{\cal J}_{52}} & = &
2 {\cal J}_{52} -{\cal J}_{12} {\cal J}_{54} - {\cal J}_{14} {\cal J}_{52},
\nonumber \\
\widetilde{{\cal J}_{53}} & = &
{\cal J}_{53} - {\cal J}_{13} {\cal J}_{54} -{\cal J}_{13} {\cal H}_{\bar{2}} {\cal J}_{52}
- {\cal J}_{43} {\cal J}_{52} + k \bar{{\cal D}} {\cal J}_{13} {\cal J}_{52}, \nonumber \\
\widetilde{{\cal J}_{35}} & = & {\cal J}_{35} + {\cal H}_{2} {\cal J}_{15} {\cal J}_{32} -
{\cal J}_{15} {\cal J}_{34} - k {\cal J}_{15} {\cal D} {\cal J}_{32} -
{\cal J}_{45} {\cal J}_{32}, \nonumber \\
\widetilde{{\cal H}_{2}} & = & {\cal H}_{2}, \nonumber \\
\widetilde{{\cal H}_{\bar{2}}} & = & {\cal H}_{\bar{1}} + {\cal H}_{\bar{2}}, \nonumber \\
\widetilde{{\cal J}_{25}} & = &
{\cal J}_{13} {\cal J}_{35} {\cal H}_{\bar{2}} + {\cal J}_{14} {\cal J}_{25} -
{\cal J}_{15} {\cal J}_{24} +
k {\cal J}_{15} {\cal D} {\cal H}_{\bar{1}} + {\cal J}_{23} {\cal J}_{15} {\cal J}_{32} +
{\cal J}_{35} {\cal J}_{43} \nonumber \\
& & +
{\cal J}_{45} {\cal H}_{\bar{1}} - k \bar{{\cal D}} {\cal J}_{13} {\cal J}_{35} -
k \bar{{\cal D}} {\cal J}_{14} {\cal J}_{45} - k^2 {\cal J}_{14}' {\cal J}_{15}, \nonumber \\
\widetilde{{\cal J}_{51}} & = &
{\cal H}_{1} {\cal J}_{54} - {\cal H}_{2} {\cal J}_{32} {\cal J}_{53} -
{\cal J}_{23} {\cal J}_{32} {\cal J}_{52} +
{\cal J}_{24} {\cal J}_{52} + {\cal J}_{34} {\cal J}_{53} + {\cal J}_{42} {\cal J}_{51}
\nonumber \\
& & -
k {\cal D} {\cal H}_{\bar{1}} {\cal J}_{52} + k {\cal D} {\cal J}_{32} {\cal J}_{53} + k {\cal D} {\cal J}_{42}
{\cal J}_{54} +
k^2 {\cal J}_{12}' {\cal J}_{54} + k^2 {\cal J}_{14}' {\cal J}_{52}, \nonumber \\
& & - k \bar{{\cal D}} {\cal J}_{12} {\cal J}_{51} \nonumber \\
\widetilde{{\cal J}_{23}} & = &
k {\cal D} {\cal J}_{43} + {\cal H}_{2} {\cal J}_{43} - k
{\cal J}_{12} \bar{{\cal D}} {\cal J}_{23} +
{\cal J}_{12} {\cal J}_{23} {\cal H}_{\bar{2}} - k {\cal J}_{13} {\cal D} {\cal H}_{\bar{2}} \nonumber \\
& & + {\cal J}_{23} {\cal J}_{14} -
k \bar{{\cal D}} {\cal H}_{2} {\cal J}_{13} + k \bar{{\cal D}} {\cal J}_{12} {\cal J}_{23} + k^2
{\cal J}_{13}', \nonumber \\
\widetilde{{\cal J}_{41}} & = &
{\cal J}_{24} + k \bar{{\cal D}} {\cal H}_{1} + k^3 \bar{{\cal D}} {\cal J}_{12}' -
k {\cal D} {\cal H}_{\bar{1}} -
{\cal J}_{12} {\cal J}_{21} \nonumber \\
& & - {\cal J}_{13} {\cal J}_{31} + {\cal J}_{14} {\cal J}_{41} -
{\cal J}_{23} {\cal J}_{32} +
k^2 {\cal J}_{14}' - k^2 {\cal J}_{42}', \nonumber \\
\widetilde{{\cal J}_{31}} & = &
k \bar{{\cal D}} {\cal J}_{34} + {\cal J}_{31} {\cal J}_{42} - {\cal J}_{32} {\cal J}_{41}-
{\cal J}_{34} {\cal H}_{\bar{2}} -
k \bar{{\cal D}} {\cal H}_{2} {\cal J}_{32} \nonumber \\
& & - k \bar{{\cal D}} {\cal J}_{12} {\cal J}_{31} - k {\cal D} {\cal H}_{\bar{2}} {\cal J}_{
32} - k^2 {\cal J}_{32}'.
\end{eqnarray}
The whole set of irreducibility constraints for surviving supercurrents
is as follows
\begin{eqnarray}
& & {\cal D} {\cal H}_{2} = 0, \; \bar{{\cal D}} {\cal H}_{\bar{2}} = 0, \;
\bar{{\cal D}} {\cal J}_{53} = 0, \nonumber \\
& & \left( {\cal D} - \frac{1}{k} {\cal H}_{2} \right)
{\cal J}_{35} = 0, \;
\left( {\cal D} + \frac{1}{k} {\cal H}_{2} \right)
{\cal J}_{23} = 0, \nonumber \\
& & {\cal D} {\cal J}_{25} - \frac{1}{k} {\cal J}_{23}
{\cal J}_{35} = 0, \;
\left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{2}} \right)
{\cal J}_{31} = 0, \nonumber \\
& & \left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{2}} \right)
{\cal J}_{51}+ \frac{1}{k} \left(
{\cal J}_{21} {\cal J}_{52}+
{\cal J}_{31} {\cal J}_{53}+ {\cal J}_{41} {\cal J}_{54} \right) = 0.
\label{eq:5cons}
\end{eqnarray}
The stress tensor has the following form:
\begin{eqnarray}
{\cal T}=-\frac{1}{k} \left[ {\cal H}_{2} {\cal H}_{\bar{2}}+
{\cal J}_{41}+
{\cal J}_{15} {\cal J}_{51}+ {\cal J}_{25}
{\cal J}_{52}+
{\cal J}_{35} {\cal J}_{53}+
k {\cal D} {\cal J}_{15} ( k \bar{{\cal D}} {\cal J}_{52}- {\cal H}_{\bar{2}}
{\cal J}_{52} ) \right]
\label{eq:str}
\end{eqnarray}
and possesses the central charge $-2 k$.
On the constraints shell the Sugawara $N=2$ stress tensor coincides with
${\cal T}$. With respect to ${\cal T}$, the following combinations of supercurrents
\begin{eqnarray}
{\cal J}_{15}, \;\;
{\cal J}_{52}, \;\;
{\cal J}_{53}, \;\;
{\cal J}_{23}, \;\;
{\cal H}_{2}, \;\;
{\cal H}_{\bar{2}}, \;\;
{\cal J}_{35}, \;\;
{\cal J}_{31}, \;\;
{\cal J}_{25}-\frac{k^2}{2} ( [ {\cal D}, \bar{{\cal D}} ] {\cal J}_{15}-
{\cal J}_{15}'), \nonumber \\
{\cal J}_{51}-\frac{k^2}{2} [ {\cal D}, \bar{{\cal D}} ] {\cal J}_{52}-
\frac{k^2}{2} {\cal J}_{52}'+k {\cal D} {\cal H}_{\bar{2}} {\cal J}_{52}+
k {\cal J}_{52} {\cal T}
\label{eq:primar}
\end{eqnarray}
are superprimary.
It is straightforward to derive the complete set of SOPEs
between the above supercurrents $
{\cal J}_{\alpha \beta}$'s (\ref{eq:primar}). The $N=2$ stress tensor
(\ref{eq:str}) entering into this $N=2$ SCA is {\it unconstrained}
since the linear term ${\cal J}_{41}$ in
(\ref{eq:str}) is unconstrained. We do not give here the SOPEs between
the surviving supercurrents because these are
very complicated due to the presence of dimension zero supercurrents
${\cal J}_{15}, {\cal J}_{52}$. Let us only point out
that in the present case one cannot decouple
two fields of dimension $0$ after passing to the component form of
the superalgebra.
In the next Subsection we will show that the above
unpleasant features of SCA under consideration disappear after
the appropriate secondary hamiltonian reduction of it.
The resulting SCA does not contain any spin $0$ supercurrents; all the
involved supercurrents are superprimary with respect to the corresponding
$N=2$ stress tensor. This reduction is accomplished by
adding two more constraints to the set \p{eq:consn5} and so corresponds
to imposing some seven constraints on the original supermatrix of
$N=2$ $sl(3|2)^{(1)}$ affine supercurrents.
\subsection{\bf $N=2$ $u(2|1)$ SCA}
In this Subsection we show that there exists a natural reduction
of the second of extended $N=2$ SCAs considered in the previous
Subsection, such that it yields a $N=2$ extension of the $u(2|1)$ SCA of
Ref. \cite{DTH}.
The $u(2|1)$ SCA is some graded version of the $u(3)$
KB SCA and is generated by $16$ component currents,
the number of bosonic and fermionic ones being the same.
The spins of them are greater than $1/2$. The details of this algebra
will be given later, the only point we wish to mention at once is
that there is no standard supersymmetry subalgebra in this SCA.
Anticipating our results, the $N=2$ supersymmetric extension of this SCA,
$N=2$ $u(2|1)$, contains four extra spin $1/2$ currents: two of them are
bosonic, others fermionic.
This current content immediately implies that the number of
the hamiltonian reduction constraints should be seven. One could start
directly from $N=2$ $sl(3|2)^{(1)}$ current algebra, i.e. make use of the
primary hamiltonian reduction procedure. However, it is simpler to
deduce the same results in an equivalent way, applying a secondary
reduction to the extended $N=2$ SCA described in the end of
previous Subsection.
Thus we start with the same choice of splitting parameters
(\ref{eq:new5para}) and wish to strengthen the set of constraints
\p{eq:consn5} by adding two more ones. A natural desire
is to get rid of the unwanted spin $0$ supercurrents, viz.
${\cal J}_{15}$, ${\cal J}_{25}$ (see Table 9). It turns out that they both are
eliminated by enforcing the constraint
\begin{eqnarray}
{\cal J}_{15} = 0.
\label{eq:constru21}
\end{eqnarray}
Then we can gauge away ${\cal J}_{52}$ using the gauge
transformation generated by this new constraint:
\begin{eqnarray}
{\cal J}_{52} = 0.
\label{eq:fixu21}
\end{eqnarray}
We also note that \p{eq:constru21}, via the nonlinear
constraints (\ref{eq:cons}), automatically implies
\begin{eqnarray}
{\cal J}_{45} = 0\;.
\end{eqnarray}
So the final supermatrix of constraints is given by
\begin{eqnarray}
{\cal J}_{mn}^{constr}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & 0 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right).
\label{eq:consu2}
\end{eqnarray}
As in previous examples, we list the constrained, gauge fixed and
surviving supercurrents in Tables 10 and 11.
\begin{eqnarray}
\begin{array}{cccccccc}
\mbox{Table 10} \nonumber \\
\hline
u(1) & 1 & 2 & 0 & 0 & 1 & 2 & 3 \nonumber \\
\mbox{dim} & -\frac{1}{2} & 0 & 0 & 0 & \frac{1}{2} & 0 &
\frac{1}{2} \nonumber \\
\hline
\mbox{constr. scs} & {\cal J}_{12}^F & {\cal J}_{13}^F & {\cal J}_{42}^B &
{\cal J}_{14}^B & {\cal J}_{43}^B & {\cal J}_{15}^B & {\cal J}_{45}^F \nonumber \\
\hline
\hline
\mbox{g.f. scs} & {\cal J}_{24}^B &{\cal J}_{34}^B &
{\cal H}_{\bar{1}}^F & {\cal H}_{1}^F &
{\cal J}_{32}^F & {\cal J}_{52}^B & {\cal J}_{54}^F \nonumber \\ \hline
\mbox{dim} & 1 & \frac{1}{2} & \frac{1}{2} & \frac{1}{2} & 0 & 0 &
\frac{1}{2} \nonumber \\
u(1) & 0 & -1 & -1 & 1 & -2 & -2 & -3 \nonumber \\
\hline
\end{array}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{cccccccccc||c}
\mbox{Table 11} \nonumber \\
\hline
\mbox{surv. scs} & {\cal J}_{53}^B & {\cal J}_{35}^B &
{\cal H}_{2}^F & {\cal H}_{\bar{2}}^F & {\cal J}_{25}^B & {\cal J}_{51}^B & {\cal J}_{23}^F
& {\cal J}_{41}^B &
{\cal J}_{31}^F & {\cal J}_{21}^F \nonumber \\ \hline
\mbox{dim} & \frac{1}{2} & \frac{1}{2} & \frac{1}{2} &
\frac{1}{2} & 1 & 1 & 1 & 1 & 1 &
\frac{3}{2} \nonumber \\
u(1) & -1 & 1 & 1 & -1 & 2 & -2 & 2 & 0 & -2 & -1 \nonumber \\
\hline
\end{array}
\end{eqnarray}
After substituting (\ref{eq:constru21}), (\ref{eq:fixu21})
into (\ref{eq:dsds}), the relevant ${\cal J}_{mn}^{DS}$ takes the following form
\begin{eqnarray}
{\cal J}_{mn}^{DS}=
\left( \begin{array}{ccccc}
0 & 0 & 0 & 1 & 0 \\
k \bar{{\cal D}} {\cal J}_{41} & {\cal H}_{\bar{2}} & {\cal J}_{23} & 0 & {\cal J}_{25} \\
{\cal J}_{31} & 0 & {\cal H}_{2} & 0 & {\cal J}_{35} \\
{\cal J}_{41} & 1 & 0 & 0 & 0 \\
{\cal J}_{51} & 0 & {\cal J}_{53} & 0 & {\cal H}_{2}+{\cal H}_{\bar{2}}
\end{array} \right).
\end{eqnarray}
All the elementary supercurrents here, except for ${\cal J}_{41}$,
are still subjected to the constraints which are obtained
by substituting (\ref{eq:constru21}), (\ref{eq:fixu21})
into (\ref{eq:5cons}):
\begin{eqnarray}
{\cal D} {{\cal H}_{2}} = 0, & \;\; &
\bar{{\cal D}} {\cal H}_{\bar{2}} = 0, \nonumber \\
\left( {\cal D} - \frac{1}{k} {{\cal H}_{2}} \right)
{\cal J}_{35} = 0, & \;\; &
\bar{{\cal D}} {\cal J}_{53} = 0, \nonumber \\
\left( {\cal D} + \frac{1}{k} {{\cal H}_{2}} \right)
{{\cal J}_{23}} = 0, & \;\; &
\left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{2}} \right)
{\cal J}_{31} = 0, \nonumber \\
{\cal D} {{\cal J}_{25}} - \frac{1}{k} {{\cal J}_{23}}
{\cal J}_{35} = 0, & \;\; &
\left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{2}} \right)
{\cal J}_{51}-\frac{1}{k} {\cal J}_{31} {\cal J}_{53}
= 0.
\label{eq:ucons}
\end{eqnarray}
Using the same techniques as before, we get the following expressions for
gauge invariant supercurrents in terms of the original ones (forming the
previous SCA with five constraints)
\begin{eqnarray}
\widetilde{{\cal J}_{35}} & = & {\cal J}_{35}, \nonumber \\
\widetilde{{\cal J}_{53}} & = & {\cal J}_{53}, \nonumber \\
\widetilde{{\cal H}_{\bar{2}}} & = & {\cal H}_{\bar{2}}, \nonumber \\
\widetilde{{\cal H}_{2}} & = & {\cal H}_{2}, \nonumber \\
\widetilde{{\cal J}_{25}} & = & {\cal J}_{25} - \frac{k^2}{2} [{\cal D}, \bar{{\cal D}} ]
{\cal J}_{15} - k {\cal J}_{15} {\cal D} {\cal H}_{\bar{2}} +
{\cal J}_{15}
{\cal J}_{35} {\cal J}_{53} -
k {\cal D} {\cal J}_{15} {\cal H}_{\bar{2}} + \frac{k^2}{2}
{\cal J}_{15}', \nonumber \\
\widetilde{{\cal J}_{23}} & = & {\cal J}_{23} + {\cal H}_{2}
{\cal J}_{15}
{\cal J}_{53} + k {\cal J}_{15} {\cal D} {\cal J}_{53} + k
{\cal D} {\cal J}_{15} {\cal J}_{53}, \nonumber \\
\widetilde{{\cal J}_{41}} & = & {\cal J}_{41} + {\cal J}_{15}
{\cal J}_{51} +
{\cal J}_{25} {\cal J}_{52}, \nonumber \\
\widetilde{{\cal J}_{31}} & = & {\cal J}_{31} + k {\cal J}_{35} \bar{{\cal D}}
{\cal J}_{
52} - {\cal J}_{35} {\cal H}_{\bar{2}} {\cal J}_{52} +
k \bar{{\cal D}} {\cal J}_{35} {\cal J}_{52}, \nonumber \\
\widetilde{{\cal J}_{51}} & = & {\cal J}_{51} - \frac{k^2}{2} [{\cal D}, \bar{{\cal D}} ]
{\cal J}_{52} - k {\cal H}_{\bar{2}} {\cal D} {\cal J}_{52} +
{\cal J}_{35} {\cal J}_{52} {\cal J}_{53} -
{\cal J}_{41} {\cal J}_{52} - \frac{k^2}{2}
{\cal J}_{52}'.
\end{eqnarray}
We would like to stress that the SOPEs between ${\cal J}_{kl}$
appearing in the r.h.s. of above equations can be found by
using the SOPEs of second superalgebra presented in the Subsection $5.3$.
The $N=2$ stress tensor is given by
\begin{eqnarray}
{\cal T}=-\frac{1}{k} \left[ {\cal H}_{2} {\cal H}_{\bar{2}}+
{\cal J}_{41}+
{\cal J}_{35} {\cal J}_{53} \right]
\label{eq:u21T}
\end{eqnarray}
with central charge $-2 k$. On the shell of constraints Sugawara
$N=2$ stress tensor coincides with this stress tensor and
contains linearly supercurrent ${\cal J}_{41}$, so ${\cal T}$
(\ref{eq:u21T}) is unconstrained.
Then $N=2$ $u(2|1)$ SCA (the reason why we call it this way will be
soon clear) besides general spin $1$
supercurrent ${\cal T}$, $N=2$ stress tensor, contains the following
eight constrained $N=2$ supercurrents: spin $1/2$ ${\cal H}_{2}$,
${\cal J}_{53}$, ${\cal J}_{35}$ and ${\cal H}_{\bar{2}}$, spin $1$ ${\cal J}_{51},
{\cal J}_{25},$ ${\cal J}_{23},$ and ${\cal J}_{31}$. All these supercurrents are
superprimary with respect to ${\cal T}$. After rescaling
\begin{eqnarray}
{\cal J}_{25} \rightarrow \frac{1}{k} {\cal J}_{25}, \;\;
{\cal J}_{23} \rightarrow \frac{1}{k} {\cal J}_{23}, \;\;
{\cal J}_{31} \rightarrow \frac{1}{k} {\cal J}_{31}, \;\;
{\cal J}_{51} \rightarrow \frac{1}{k} {\cal J}_{51},
\end{eqnarray}
the rest of nonvanishing SOPEs are as follows
\vspace{2cm}
\begin{eqnarray}
{{\cal H}_{2}} (Z_{1}) {\cal H}_{\bar{2}} (Z_{2}) & = &
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \frac{k}{2} -\frac{1}{z_{12}} k, \nonumber \\
{\cal H}_{\bar{2}} (Z_{1}) {\cal J}_{35} (Z_{2}) & = &
-\frac{\theta_{12}}{z_{12}} {\cal J}_{35}, \nonumber \\
{\cal H}_{\bar{2}} (Z_{1}) {\cal J}_{53} (Z_{2}) & = &
\frac{\theta_{12}}{z_{12}} {\cal J}_{53}, \nonumber \\
{\cal J}_{35} (Z_{1}) {\cal J}_{53} (Z_{2}) & = &
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \frac{k}{2} +\frac{1}{z_{12}} k + \frac{\theta_{12}}{z_{12}} {{\cal H}_{2}} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \bar{{\cal D}} {{\cal H}_{2}},
\label{eq:u11}
\end{eqnarray}
\begin{eqnarray}
& & {{\cal H}_{2}} (Z_{1})
\left\{\begin{array}{ccl}
{{\cal J}_{23}} (Z_{2}) & = &
\frac{\bar{\theta}_{12}}{z_{12}} {{\cal J}_{23}}, \nonumber \\
{{\cal J}_{25}} (Z_{2}) & = &
\frac{\bar{\theta}_{12}}{z_{12}} {{\cal J}_{25}}, \nonumber \\
{\cal J}_{31} (Z_{2}) & = &
-\frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{31}, \nonumber \\
{\cal J}_{51} (Z_{2}) & = &
-\frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{51}, \nonumber \\
\end{array} \right. \nonumber \\
& & {\cal H}_{\bar{2}} (Z_{1})
\left\{\begin{array}{ccl}
{{\cal J}_{23}} (Z_{2}) & = &
\frac{\theta_{12}}{z_{12}} {{\cal J}_{23}}, \nonumber \\
{\cal J}_{31} (Z_{2}) & = &
-\frac{\theta_{12}}{z_{12}} {\cal J}_{31}, \nonumber \\
\end{array} \right. \nonumber \\
& & {\cal J}_{35} (Z_{1})
\left\{\begin{array}{ccl}
{{\cal J}_{23}} (Z_{2}) & = &
-\frac{\bar{\theta}_{12}}{z_{12}} {{\cal J}_{25}} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} \left[
{{\cal H}_{2}} {{\cal J}_{25}}+{\cal J}_{35}
{{\cal J}_{23}} \right], \nonumber \\
{{\cal J}_{25}} (Z_{2}) & = &
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} {\cal J}_{35} {{\cal J}_{25}}, \nonumber \\
{\cal J}_{31} (Z_{2}) & =&
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} {\cal J}_{35} {\cal J}_{31},
\nonumber \\
{\cal J}_{51} (Z_{2}) & = &
-\frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{31} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}}
\frac{1}{k} \left[ {{\cal H}_{2}} {\cal J}_{31} -
{\cal J}_{35} {\cal J}_{51} \right], \nonumber \\
\end{array} \right. \nonumber \\
& & {\cal J}_{53} (Z_{1})
\left\{\begin{array}{ccl}
{{\cal J}_{25}} (Z_{2}) & = &
\frac{\theta_{12}}{z_{12}} {{\cal J}_{23}}, \nonumber \\
{\cal J}_{31} (Z_{2}) & = &
-\frac{\theta_{12}}{z_{12}} {\cal J}_{51}, \nonumber \\
\end{array} \right.
\end{eqnarray}
\begin{eqnarray}
{{\cal J}_{23}} (Z_{1}) {{\cal J}_{25}} (Z_{2}) & = &
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} {{\cal J}_{23}} {{\cal J}_{25}},
\nonumber \\
{{\cal J}_{23}} (Z_{1}) {\cal J}_{31} (Z_{2}) & = &
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^3} k-
\frac{1}{z_{12}^{2}} k + \frac{\bar{\theta}_{12}}{z_{12}^2} {\cal H}_{\bar{2}} +
\frac{{\theta}_{12}}{z_{12}^2} {{\cal H}_{2}} \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[-\frac{1}{2} {\cal T} + \frac{3}{2} \bar{{\cal D}} {{\cal H}_{2}} +
\frac{1}{2} {\cal D} {\cal H}_{\bar{2}} + \frac{1}{2 k}
{{\cal H}_{2}}
{\cal H}_{\bar{2}} -
\frac{1}{k} {\cal J}_{35} {\cal J}_{53} \right] \nonumber \\
& & +
\frac{1}{z_{12}} \left[ {\cal T}- {\cal D} {\cal H}_{\bar{2}} +\frac{1}{k}
{{\cal H}_{2}} {\cal H}_{\bar{2}}+
\frac{2}{k} {\cal J}_{35}
{\cal J}_{53} - \bar{{\cal D}} {{\cal H}_{2}} \right] \nonumber \\
& & + \frac{\bar{\theta}_{12}}{z_{12}} \left[ \bar{{\cal D}} {\cal T}-\frac{1}{k} {\cal H}_{\bar{2}} {\cal T}+
\frac{1}{k} {\cal H}_{\bar{2}}
{\cal D} {\cal H}_{\bar{2}}-\frac{2}{k^2}
{\cal J}_{35} {\cal H}_{\bar{2}} {\cal J}_{53}+
\frac{2}{k} \bar{{\cal D}} {{\cal H}_{2}}
{\cal H}_{\bar{2}} \right. \nonumber \\
& & \left. + \frac{2}{k} \bar{{\cal D}} {\cal J}_{35} {\cal J}_{53}+
{\cal H}_{\bar{2}}'\right] \nonumber \\
& & +
\frac{\theta_{12}}{z_{12}} \left[ -\frac{1}{k} {{\cal H}_{2}} {\cal T}+\frac{1}{k}
{{\cal H}_{2}} \bar{{\cal D}} {{\cal H}_{2}} +\frac{1}{k}
{{\cal H}_{2}}
{\cal D} {\cal H}_{\bar{2}}-\frac{2}{k^2}
{{\cal H}_{2}} {\cal J}_{35} {\cal J}_{53}+
{{\cal H}_{2}}' \right]
\nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[
\bar{{\cal D}} {{\cal H}_{2}}'+\frac{1}{k} {{\cal H}_{2}} \bar{{\cal D}} {\cal T}-
\frac{1}{k^2} {{\cal H}_{2}} {\cal H}_{\bar{2}} {\cal T}+
\frac{1}{k^2} {{\cal H}_{2}} {\cal H}_{\bar{2}} {\cal D}
{\cal H}_{\bar{2}} \right. \nonumber \\
& & - \frac{2}{k^3} {{\cal H}_{2}} {\cal J}_{35}
{\cal H}_{\bar{2}} {\cal J}_{53}+
\frac{1}{k^2} {{\cal H}_{2}} \bar{{\cal D}} {{\cal H}_{2}}
{\cal H}_{\bar{2}}+\frac{2}{k^2} {{\cal H}_{2}} \bar{{\cal D}}
{\cal J}_{35}
{\cal J}_{53}+
\frac{1}{k} {{\cal J}_{23}} {\cal J}_{31} \nonumber \\
& & - \frac{1}{k} \bar{{\cal D}} {{\cal H}_{2}} {\cal T}+\frac{1}{k}
\bar{{\cal D}} {{\cal H}_{2}} \bar{{\cal D}} {\cal H}_{2}+\frac{1}{k} \bar{{\cal D}} {{\cal H}_{2}} {\cal D} {\cal H}_{\bar{2}}-
\frac{2}{k^2} \bar{{\cal D}} {{\cal H}_{2}} {\cal J}_{35}
{\cal J}_{53} \nonumber \\
& & \left. +
\frac{1}{k} \left( {{\cal H}_{2}}
{\cal H}_{\bar{2}} \right)' \right], \nonumber \\
{{\cal J}_{23}} (Z_{1}) {\cal J}_{51} (Z_{2}) & = &
- \frac{\bar{\theta}_{12}}{z_{12}^2} {\cal J}_{53} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[- \frac{1}{2}
{\cal D} {\cal J}_{53} -
\frac{3}{2 k} {{\cal H}_{2}} {\cal J}_{53} \right] \nonumber \\
& & +
\frac{1}{z_{12}} \left[ {\cal D} {\cal J}_{53}+\frac{1}{k}
{{\cal H}_{2}} {\cal J}_{53} \right] \nonumber \\
& & +
\frac{\bar{\theta}_{12}}{z_{12}} \left[-\frac{1}{k} {\cal H}_{\bar{2}} {\cal D} {\cal J}_{53}+
\frac{2}{k^2} {{\cal H}_{2}}
{\cal H}_{\bar{2}} {\cal J}_{53}+
\frac{2}{k^2} {\cal J}_{35} {\cal J}_{53} {\cal J}_{53}+
\frac{1}{k} {\cal T} {\cal J}_{53}
\right. \nonumber \\
& & \left. - \frac{1}{k} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{53}-
{\cal J}_{53}' \right] - \frac{\theta_{12}}{z_{12}} \frac{1}{k}
{{\cal H}_{2}} {\cal D} {\cal J}_{53} \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[-\frac{1}{k^2} {{\cal H}_{2}}
{\cal H}_{\bar{2}} {\cal D} {\cal J}_{53}+\frac{2}{k^3}
{{\cal H}_{2}} {\cal J}_{35}
{\cal J}_{53} {\cal J}_{53}
+ \frac{1}{k^2} {{\cal H}_{2}} {\cal T} {\cal J}_{53} \right. \nonumber \\
& & - \frac{1}{k^2}
{{\cal H}_{2}} \bar{{\cal D}} {{\cal H}_{2}} {\cal J}_{53} -
\frac{1}{k^2} {{\cal H}_{2}} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{
53}
-\frac{1}{k} \left( {{\cal H}_{2}} {\cal J}_{53} \right)'+
\frac{1}{k} {{\cal J}_{23}} {\cal J}_{51} \nonumber \\
& & \left. -
\frac{1}{k} \bar{{\cal D}} {{\cal H}_{2}} {\cal D}
{\cal J}_{53} \right] \nonumber \\
{{\cal J}_{25}} (Z_{1}) {\cal J}_{31} (Z_{2}) & = &
-\frac{{\theta}_{12}}{z_{12}^2} {\cal J}_{35} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[ -\frac{3}{2} \bar{{\cal D}}
{\cal J}_{35} +
\frac{1}{k} {\cal J}_{35} {\cal H}_{\bar{2}} \right] \nonumber \\
& & +
\frac{1}{z_{12}} \bar{{\cal D}} {\cal J}_{35}+ \frac{\bar{\theta}_{12}}{z_{12}} \frac{1}{k} \bar{{\cal D}}
{\cal J}_{35} {\cal H}_{\bar{2}} \nonumber \\
& & + \frac{\theta_{12}}{z_{12}} \left[ \frac{1}{k^2} {{\cal H}_{2}}
{\cal J}_{35} {\cal H}_{\bar{2}}+\frac{1}{k}
{\cal J}_{35} {\cal T}- \frac{1}{k} {\cal J}_{35} {\cal D}
{\cal H}_{\bar{2}}+
\frac{2}{k^2} {\cal J}_{35} {\cal J}_{35}
{\cal J}_{53} \right. \nonumber \\
& & \left. - \frac{1}{k} \bar{{\cal D}} {{\cal H}_{2}} {\cal J}_{35}-
{\cal J}_{35}' \right] \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[-
\bar{{\cal D}} {\cal J}_{35}'-\frac{1}{k^2} {{\cal H}_{2}} \bar{{\cal D}}
{\cal J}_{35} {\cal H}_{\bar{2}}+\frac{1}{k}
{\cal J}_{35}
\bar{{\cal D}} {\cal T}-\frac{1}{k^2} {\cal J}_{35}
{\cal H}_{\bar{2}} {\cal T} \right. \nonumber \\
& & + \frac{1}{k^2} {\cal J}_{35} {\cal H}_{\bar{2}} {\cal D}
{\cal H}_{\bar{2}}-\frac{2}{k^3}
{\cal J}_{35} {\cal J}_{35}
{\cal H}_{\bar{2}} {\cal J}_{53}+\frac{4}{k^2}
{\cal J}_{35} \bar{{\cal D}} {\cal J}_{35} {\cal J}_{53}+
\frac{1}{k} \left( {\cal J}_{35} {\cal H}_{\bar{2}} \right)'
\nonumber \\
& & \left.
- \frac{1}{k} \bar{{\cal D}} {{\cal H}_{2}} \bar{{\cal D}} {\cal J}_{35}+\frac{2}{k^2}
\bar{{\cal D}} {{\cal H}_{2}} {\cal J}_{35} {\cal H}_{\bar{2}}+
\frac{1}{k} \bar{{\cal D}} {\cal J}_{35} {\cal T}
-\frac{1}{k}
\bar{{\cal D}} {\cal J}_{35} {\cal D} {\cal H}_{\bar{2}} \right],
\nonumber \\
{{\cal J}_{25}} (Z_{1}) {\cal J}_{51} (Z_{2}) & = &
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^3} k
-\frac{1}{z_{12}^2} k + \frac{\bar{\theta}_{12}}{z_{12}^2} {\cal H}_{\bar{2}} \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[ -\frac{1}{2} {\cal T} +
\frac{1}{2} {\cal D} {\cal H}_{\bar{2}} - \frac{1}{2 k}
{{\cal H}_{2}} {\cal H}_{\bar{2}}-
\frac{2}{k} {\cal J}_{35} {\cal J}_{53} \right] \nonumber \\
& & +
\frac{1}{z_{12}} \left[ {\cal T} - {\cal D} {\cal H}_{\bar{2}}+\frac{1}{k}
{{\cal H}_{2}} {\cal H}_{\bar{2}}+
\frac{2}{k} {\cal J}_{35}
{\cal J}_{53} \right] \nonumber \\
& & + \frac{\bar{\theta}_{12}}{z_{12}} \left[ \bar{{\cal D}} {\cal T}-\frac{1}{k} {\cal H}_{\bar{2}} {\cal T}+
\frac{1}{k} {\cal H}_{\bar{2}}
{\cal D} {\cal H}_{\bar{2}}-
\frac{2}{k^2} {\cal J}_{35} {\cal H}_{\bar{2}}
{\cal J}_{53} \right. \nonumber \\
& & \left. + \frac{1}{k} \bar{{\cal D}} {{\cal H}_{2}} {\cal H}_{\bar{2}}+
\frac{1}{k} \bar{{\cal D}} {\cal J}_{35} {\cal J}_{53}+
{\cal H}_{\bar{2}}' \right] + \nonumber \\
& & \frac{\theta_{12}}{z_{12}} \left[ \frac{1}{k^2} {{\cal H}_{2}} {\cal J}_{35}
{\cal J}_{53}
+\frac{1}{k} {\cal J}_{35} {\cal D} {\cal J}_{53} \right] \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[\frac{2}{k^3} {{\cal H}_{2}} {\cal J}_{35}
{\cal H}_{\bar{2}}
{\cal J}_{53}-
\frac{1}{k^2} {{\cal H}_{2}} \bar{{\cal D}} {\cal J}_{35}
{\cal J}_{53} + \frac{1}{k} {{\cal J}_{23}} {\cal J}_{31}
\right. \nonumber \\
& & -
\frac{1}{k^2} {\cal J}_{35} {\cal H}_{\bar{2}} {\cal D}
{\cal J}_{53}
+\frac{2}{k^3} {\cal J}_{35} {\cal J}_{35}
{\cal J}_{53} {\cal J}_{53}+ \frac{1}{k^2}
{\cal J}_{35} {\cal T} {\cal J}_{53} \nonumber \\
& & \left. + \frac{1}{k} \bar{{\cal D}} {\cal J}_{35} {\cal D} {\cal J}_{53}-
\frac{1}{k}
\left( {\cal J}_{35} {\cal J}_{53} \right)'
-\frac{1}{k^2} {\cal J}_{35} {\cal D} {\cal H}_{\bar{2}} {\cal J}_{53}
\right], \nonumber \\
\end{eqnarray}
It can be checked that our algebra satisfies all the Jacobi identities.
The supercurrents ${\cal H}_{2}, {\cal H}_{\bar{2}}, {\cal J}_{35}, {\cal J}_{53}$ form $N=2$
$u(1|1)^{(1)}$ current algebra as a subalgebra (their SOPEs form a closed
set as is seen from eq. (\ref{eq:u11})).
Let us show that $N=2$ $W_{3}^{(2)}$ SCA constructed in Subsection
$5.2$ using primary hamiltonian reduction can be equally obtained
via a secondary hamiltonian reduction from $N=2$ $u(2|1)$ SCA. The
existence of such a possibility follows already from the fact that
the constraints \p{eq:consu2} form a subclass of the $N=2$ $W_3^{(2)}$
constraints \p{eq:consw32}.
With respect to the new stress tensor ${\cal T}_{\mbox{new}}$
\begin{eqnarray}
{\cal T}_{\mbox{new}}= {\cal T} - 2 \bar{{\cal D}} {\cal H}_{2}
\end{eqnarray}
the spins ($u(1)$ charges) of ${\cal J}_{25}$ and ${\cal J}_{23}$
are $0(0)$. In this new basis,
we can add two extra constraints such that
\begin{eqnarray}
{\cal J}_{25}=1, \;\; {\cal J}_{23}=0\;,
\label{eq:cons25}
\end{eqnarray}
and, as usual, make use of the gauge freedom associated with
these constraints for gauging away two more supercurrents
\begin{eqnarray}
{\cal H}_{2}=0, \;\; {\cal J}_{35}=0\;.
\label{eq:cons2}
\end{eqnarray}
Using (\ref{eq:cons25}), (\ref{eq:cons2}) one can check
that (\ref{eq:ucons}) precisely reduces to
(\ref{eq:w32con}) and ${\cal J}_{mn}^{DS}$ coincides with (\ref{eq:dsw32}).
It can be easily checked that the dimensions and
$u(1)$ charges of the surviving supercurrents ${\cal J}_{53}, {\cal H}_{\bar{2}},
{\cal J}_{51}, {\cal J}_{31}$ with respect to ${\cal T}_{\mbox{new}}$
change and take the same values as in Table 5.
After finding gauge invariant supercurrents which we did not write down
explicitly, the reduced algebra becomes the algebra $N=2$ $W_{3}^{(2)}$
SCA elaborated in Subsection $5.2$.
Let us analyze in some detail the component structure of the extended
$N=2$ SCA constructed here.
After solving the constraints (\ref{eq:ucons})
for the involved supercurrents
we are left with the following set of $(10 + 10)$ currents: one Virasoro
spin $2$ stress tensor, two bosonic and four fermionic spin $3/2$ currents,
five bosonic and four fermionic spin $1$ currents, two bosonic and two
fermionic spin $1/2$ currents. For the time being we do not give the
precise relation of these currents to the components of supercurrents,
we only note that four spin $1/2$ currents appear as
the $\theta, \bar{\theta}$ independent parts of ${\cal J}_{35}, {\cal J}_{53},
{\cal H}_{\bar{2}}, {\cal H}_{\bar{2}}$. The Virasoro stress tensor,
pair of fermionic spin
$3/2$ currents and one bosonic spin $1$ current form $N=2$ SCA as
a subalgebra, while the remainder of currents are spread over
$N=2$ multiplets.
It is not too enlightening to present the OPEs
between these latter currents. For a better understanding
what we have obtained, it is more appropriate to pass, by means of
some nonlinear invertible
transformation, to another basis of the constituent currents
in which the $N=2$ multiplet structure becomes implicit but the
spin $1/2$ currents commute with all other ones and so can be factored out.
The possibility of such a factorization agrees with the general
statement of Ref. \cite{GS}. Below we give the explicit correspondence
between the modified currents (commuting with the spin $1/2$ ones) and
the initial supercurrents
\begin{eqnarray}
k {J^{2}}_{1} & = & {\cal J}_{25} \vert, \nonumber \\
k {J^{1}}_{2} & = & {\cal J}_{51} \vert, \nonumber \\
- k {J^{3}}_{1} & = & {\cal J}_{23} \vert, \nonumber \\
k {J^{1}}_{3} & = & {\cal J}_{31} \vert, \nonumber \\
k \left( {J^{1}}_{1}-{J^{2}}_{2}-{J^{3}}_{3} \right) & = &
\left(-k {\cal T}-{\cal H}_{2} {\cal H}_{\bar{2}}-
{\cal J}_{35} {\cal J}_{53} \right) \vert, \nonumber \\
{J^{2}}_{3} & = & \bar{{\cal D}} {\cal J}_{35} \vert, \nonumber \\
-{J^{3}}_{2} & = & \left( {\cal D} {\cal J}_{53}+\frac{1}{k}
{\cal J}_{53} {\cal H}_{2}
\right) \vert, \nonumber \\
{J^{3}}_{3} & = & \left( {\cal D} {\cal H}_{\bar{2}}-\frac{1}{k}
{\cal J}_{35} {\cal J}_{53}
\right) \vert, \nonumber \\
{J^{2}}_{2}+{J^{3}}_{3} & = & \bar{{\cal D}} {\cal H}_{2} \vert, \nonumber \\
T & = & \left( -\frac{1}{2k} \left[ k [ {\cal D}, \bar{{\cal D}} ] {\cal T}-
{\cal H}_{2}' {\cal H}_{\bar{2}}+
{\cal H}_{2} {\cal H}_{\bar{2}}'-
{\cal J}_{35}' {\cal J}_{53}+{\cal J}_{35}
{\cal J}_{53}' \right]
\right) \vert, \nonumber \\
-i \frac{k}{\sqrt{2}} G^{2} & = &
\left( \bar{{\cal D}} {\cal J}_{25}+\frac{1}{k} {\cal J}_{25}
{\cal H}_{\bar{2}} \right) \vert, \nonumber \\
-i \frac{k}{\sqrt{2}} {\bar{G}}_{2} & = &
{\cal D} {\cal J}_{51} \vert, \nonumber \\
-i \frac{k}{\sqrt{2}} G^{3} & = &
\left( \bar{{\cal D}} {\cal J}_{23}+\frac{1}{k} {\cal H}_{\bar{2}}
{\cal J}_{23}-\frac{1}{k}
{\cal J}_{53} {\cal J}_{25} \right) \vert, \nonumber \\
-i \frac{k}{\sqrt{2}} {\bar{G}}_{3} & = &
\left( {\cal D} {\cal J}_{31}-\frac{1}{k} {\cal H}_{2}
{\cal J}_{31}+
\frac{1}{k} {\cal J}_{51} {\cal J}_{35} \right) \vert, \nonumber \\
-i \frac{k}{\sqrt{2}} {\bar{G}}_{1} & = &
\left( -k {\cal D} {\cal T}+{\cal H}_{2} {\cal D} {\cal H}_{\bar{2}}-
\frac{1}{k} {\cal H}_{2} {\cal J}_{35} {\cal J}_{53}-
{\cal J}_{35} {\cal D} {\cal J}_{53} \right) \vert, \nonumber \\
-i \frac{k}{\sqrt{2}} G^{1} & = &
\left( -k \bar{{\cal D}} {\cal T}-\bar{{\cal D}} {\cal H}_{2} {\cal H}_{\bar{2}}-
\bar{{\cal D}} {\cal J}_{35} {\cal J}_{53}
\right) \vert,
\end{eqnarray}
where $|$ means the $\theta, \bar{\theta}$ independent part of corresponding
supercurrents. After decoupling spin $1/2$ currents the
quotient algebra includes the Virasoro stress tensor $T$, two
bosonic and four fermionic spin $3/2$ currents,
respectively, $G^{3}, {\bar{G}}_{3}$ and $G^{a}, {\bar{G}}_{a},
(a=1,2) $, five bosonic and four fermionic spin $1$ currents,
respectively ${J^{a}}_{b}, (a,b=1,2), {J^{3}}_{3}$
and ${J^{a}}_{3}, {J^{3}}_{a}, a=1,2 $.
Nine spin $1$ currents turn out to generate $u(2|1)$
current algebra\footnote{We give the relations of
$u(m|n)$ current algebra in Appendix C.}.
Spin $3/2$ currents transform under fundamental
and conjugate representaions of $u(2|1)$ for upper and lower positions
of indices. Their OPEs contain a quadratic nonlinearity in the
$u(2|1)$ currents. All the currents are primary with respect to $T$.
A simple inspection shows that this quotient algebra is none other than
the $Z_2\times Z_2$ graded extension of $u(2|1)$ current superalgebra,
$u(2|1)$ SCA \cite{DTH},
which is some graded version of the $u(3)$ KB SCA (the
precise correspondence comes out with the choice $k= -\kappa,
m = 2, n=1$ in the general formulas of \cite{DTH}). In contrast to
the original $N=2$ algebra with the spin $1/2$ currents added, the
quotient algebra does not contain the standard linear $N=2$ SCA
as a subalgebra; respectively, the $N=2$ multiplet structure of the
currents turns out to be lost. Thus we see that the adding of the
spin $1/2$ currents to the $u(2|1)$ SCA makes it possible to extend
it to some extended $N=2$ SCA, and this is why we call the latter
$N=2$ $u(2|1)$ SCA. The relation between this SCA
and its quotient by the spin $1/2$ currents strongly resembles, say,
the relation between linear $N=3$ SCA and nonlinear $so(3)$ KB
SCA \cite{GS}.
The essential difference consists, however, in that both
$N=2$ $u(2|1)$ SCA and its quotient are {\it nonlinear} algebras.
Nonetheless, we can say that the first algebra is still ``more linear''
compared to the second one, because passing to it linearizes two of
four nonlinear supersymmetries of $u(2|1)$ SCA.
Let us also remind that in the component version of hamiltonian reduction
of $sl(3|2)^{(1)}$, when we constrain both $sl(3)$ and $sl(2)$
blocks, $N=2$ $W_{3}$ or $N=2$ $W_{3}^{(2)}$ SCAs come out.
It is also known that we can obtain $u(3)$ KB SCA by imposing
constraints only on the $sl(2)$ block \cite{Ro,IM}. In terms of
component currents, $u(2|1)$ SCA corresponds to the
reduction when constraints are placed
only on the $sl(3)$ block of the $5\times 5$ $sl(3|2)^{(1)}$ supermatrix of
currents.
In the next Subsection we show that there exists another kind of
hamiltonian reduction of $N=2$ $sl(3|2)^{(1)}$ with the same number
$7$ of constraints. It yields some nonlinear extended $N=2$ SCA which by the
same reasoning as above can be called $N=2$ $u(3)$ SCA.
\subsection{\bf $N=2$ $u(3)$ SCA}
Using exactly the same arguments as given in previous Subsections, we can
continue our reduction procedure. We want to construct an $N=2$ extension
of $u(3)$ KB SCA which has $16$ component currents:
that is, $10$ bosonic currents and $6$ fermionic ones. The minimal way to
equalize the number of bosonic and fermionic currents is to add $4$
extra fermionic currents. This implies that the number of
the relevant reduction constraints should be again equal to $7$.
We choose
\begin{eqnarray}
\alpha_{1}=1, \alpha_{\bar{1}}=0, \alpha_{2}=0, \alpha_{\bar{2}}=-1
\end{eqnarray}
and list the dimensions and $u(1)$ charges of supercurrents in Tables 12
and 13.
\begin{eqnarray}
\begin{array}{cccccccc}
\mbox{Table 12} \nonumber \\
\hline
u(1) & 1 & 0 & -2 & 0 & 2 & -1 & 3 \nonumber \\
\mbox{dim} & -\frac{1}{2} & 0 & 0 & 0 & 0 & \frac{1}{2} &
\frac{1}{2} \nonumber \\
\hline
\mbox{constr. scs} & {\cal J}_{45}^F & {\cal J}_{42}^B & {\cal J}_{43}^B &
{\cal J}_{25}^B & {\cal J}_{15}^B &
{\cal J}_{23}^F & {\cal J}_{12}^F \nonumber \\
\hline
\hline
\mbox{g.f. scs} & {\cal J}_{24}^B &{\cal H}_{\bar{1}}^F & {\cal J}_{32}^F &
{\cal H}_{2}^F &
{\cal J}_{41}^B & {\cal J}_{35}^B & {\cal J}_{21}^F \nonumber \\ \hline
\mbox{dim} & 1 & \frac{1}{2} & \frac{1}{2} & \frac{1}{2} & 0 & 0 &
\frac{1}{2} \nonumber \\
u(1) & 0 & -1 & 1 &
1 & -2 & 2 & -3 \nonumber \\
\hline
\end{array}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{cccccccccc||c}
\mbox{Table 13} \nonumber \\
\hline
\mbox{surv. scs} & {\cal J}_{13}^F & {\cal J}_{31}^F & {\cal H}_{1}^F &
{\cal H}_{\bar{2}}^F & {\cal J}_{52}^B &
{\cal J}_{14}^B & {\cal J}_{34}^B & {\cal J}_{51}^B & {\cal J}_{53}^B & {\cal J}_{54}^F \nonumber \\ \hline
\mbox{dim} & \frac{1}{2} & \frac{1}{2} & \frac{1}{2} & \frac{1}{
2} & 1 & 1 & 1 & 1 & 1 & \frac{3}{2}
\nonumber \\
u(1) & 1 & -1 & 1 & -1 & 0 & 2 & 2 & -2 & -2 & -1 \nonumber \\
\hline
\end {array}
\end{eqnarray}
We impose the following reduction constraints
\begin{eqnarray}
{\cal J}_{mn}^{constr}=
\left( \begin{array}{ccccc}
\ast & 0 & \ast & \ast & 0 \\
\ast & \ast & 0 & \ast & 1 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right).
\label{eq:cou3}
\end{eqnarray}
These constraints are a subset of those we imposed in $N=2$
$W_{3}^{(2)}$ case. This implies, by the way, that
we can produce $N=2$ $W_{3}^{(2)}$ (or $N=2$ $W_{3}$) SCA by
secondary hamiltonian reduction starting with these
seven constraints and imposing two (three) more constraints.
As usual, the gauge fixing procedure goes in accord with the Table 12
and, as the result, we are left with the set of surviving currents
indicated in the Table 13.
Using the nonlinear irreducibility constraints, we may express ${\cal J}_{54}$
through the other supercurrents
\begin{eqnarray}
{\cal J}_{54}= k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{2}} \right) {\cal J}_{52}-
{\cal J}_{32} {\cal J}_{53}\;,
\end{eqnarray}
and finally arrive at the following ${\cal J}_{mn}^{DS}$
\begin{eqnarray}
{\cal J}_{mn}^{DS}=
\left( \begin{array}{ccccc}
0 & 0 & {\cal J}_{13} & {\cal J}_{14} & 0 \\
0 & {\cal H}_{\bar{2}}+{\cal H}_{1} & 0 & 0 & 1 \\
{\cal J}_{31} & 0 & 0 & {\cal J}_{34} & 0 \\
0 & 1 & 0 & {\cal H}_{1} & 0 \\
{\cal J}_{51} & {\cal J}_{52} & {\cal J}_{53} &
k \left( \bar{{\cal D}} -\frac{1}{k} {\cal H}_{\bar{2}} \right) {\cal J}_{52}
& {\cal H}_{\bar{2}}
\end{array} \right).
\end{eqnarray}
The remnants of the irreducibility constraints read
\vspace{3cm}
\begin{eqnarray}
{\cal D} {\cal H}_{1} = 0, & \;\; &
\bar{{\cal D}} {\cal H}_{\bar{2}} = 0, \nonumber \\
\left( {\cal D} + \frac{1}{k} {\cal H}_{1} \right)
{\cal J}_{13} = 0, & \;\; &
\left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{2}} \right)
{\cal J}_{31} = 0, \nonumber \\
\bar{{\cal D}} {\cal J}_{53} = 0, & \;\; &
\left( {\cal D} - \frac{1}{k} {\cal H}_{1} \right)
{\cal J}_{34} = 0, \nonumber \\
{\cal D} {\cal J}_{14} - \frac{1}{k} {\cal J}_{13}
{\cal J}_{34} = 0, & \;\; &
\left( \bar{{\cal D}} - \frac{1}{k} {\cal H}_{\bar{2}} \right)
{\cal J}_{51}-\frac{1}{k} {\cal J}_{31} {\cal J}_{53}
= 0.
\label{eq:u3cons}
\end{eqnarray}
The computation of gauge invariant supercurrents is not very hard
due to the absence of dimension $0$ supercurrents among the surviving
currents. We give the results without entering into details
\begin{eqnarray}
\widetilde{{\cal J}_{13}} & = &
3 {\cal J}_{13} - {\cal J}_{13} {\cal J}_{25} -
{\cal J}_{13} {\cal J}_{42} - k {\cal J}_{13} \bar{{\cal D}} {\cal J}_{45}, \nonumber \\
\widetilde{{\cal J}_{31}} & = &
-{\cal J}_{31} + {\cal J}_{25} {\cal J}_{31} + {\cal J}_{31} {\cal J}_{42} +
k \bar{{\cal D}} {\cal J}_{45} {\cal J}_{31}, \nonumber \\
\widetilde{{\cal H}_{1}} & = &
{\cal H}_{1} + {\cal H}_{2} + k {\cal D} {\cal J}_{42} - k
{\cal H}_{1} \bar{{\cal D}} {\cal J}_{45} + k \bar{{\cal D}} {\cal H}_{1} {\cal J}_{45} -
k^2 {\cal J}_{45}', \nonumber \\
\widetilde{{\cal H}_{\bar{2}}} & = &
{\cal H}_{\bar{1}} + {\cal H}_{\bar{2}} + k \bar{{\cal D}} {\cal J}_{25}, \nonumber \\
\widetilde{{\cal J}_{52}} & = &
{\cal J}_{24} + k \bar{{\cal D}} {\cal H}_{2} - k {\cal D} {\cal H}_{\bar{1}} +
{\cal H}_{1} {\cal H}_{\bar{1}} + {\cal H}_{2} {\cal H}_{\bar{2}} +
{\cal J}_{14} {\cal J}_{41} + \nonumber \\
& & {\cal J}_{15} {\cal J}_{51} + {\cal J}_{25} {\cal J}_{52} + {\cal J}_{34}
{\cal J}_{43} +
{\cal J}_{35} {\cal J}_{53} + {\cal J}_{45} {\cal J}_{54}, \nonumber \\
\widetilde{{\cal J}_{14}} & = &
{\cal J}_{14} - k \bar{{\cal D}} {\cal J}_{12} - k {\cal H}_{1} \bar{{\cal D}} {\cal J}_{
15} - {\cal J}_{13} {\cal J}_{32} -
{\cal J}_{13} {\cal J}_{15} {\cal J}_{31} + \nonumber \\
& & {\cal J}_{13} {\cal J}_{35} {\cal H}_{\bar{2}} -
k {\cal J}_{14} \bar{{\cal D}} {\cal J}_{45} + k \bar{{\cal D}} {\cal H}_{1} {\cal J}_{15} - k
\bar{{\cal D}} {\cal J}_{13} {\cal J}_{35} -
k \bar{{\cal D}} {\cal J}_{14} {\cal J}_{45} - k^2 {\cal J}_{15}', \nonumber \\
\widetilde{{\cal J}_{34}} & = &
-{\cal J}_{34} + k {\cal D} {\cal J}_{32} - {\cal H}_{1} {\cal J}_{32} -
k {\cal H}_{1} \bar{{\cal D}} {\cal J}_{35} -
2 {\cal H}_{1} {\cal J}_{15} {\cal J}_{31} + \nonumber \\
& & {\cal H}_{1} {\cal J}_{35} {\cal H}_{\bar{2}} +
{\cal J}_{12} {\cal J}_{31} +
{\cal J}_{13} {\cal J}_{35} {\cal J}_{31} + {\cal J}_{14} {\cal J}_{45} {\cal J}_{31} +
k {\cal J}_{15} {\cal D} {\cal J}_{31} + \nonumber \\
& & {\cal J}_{25} {\cal J}_{34} +{\cal J}_{34} {\cal J}_{42} -
{\cal J}_{34} {\cal J}_{45} {\cal H}_{\bar{2}} - k {\cal J}_{35} {\cal D} {\cal H}_{\bar{2}} - k
\bar{{\cal D}} {\cal J}_{34} {\cal J}_{45} -
k^2 {\cal J}_{35}', \nonumber \\
\widetilde{{\cal J}_{51}} & = &
-{\cal J}_{51} + k {\cal D} {\cal J}_{21} -{\cal H}_{1} {\cal J}_{21} +
{\cal H}_{1} {\cal H}_{\bar{2}} {\cal J}_{41} -
{\cal H}_{1} {\cal J}_{31} {\cal J}_{43} + \nonumber \\
& & k {\cal H}_{\bar{2}} {\cal D} {\cal J}_{41} -
{\cal J}_{13} {\cal J}_{31} {\cal J}_{41} -
{\cal J}_{23} {\cal J}_{31} + {\cal J}_{25} {\cal J}_{51} - {\cal J}_{41} {\cal J}_{52} + {\cal J}_{42}
{\cal J}_{51} + \nonumber \\
& &
k \bar{{\cal D}} {\cal H}_{1} {\cal J}_{41} + k \bar{{\cal D}} {\cal J}_{45} {\cal J}_{51} + k {\cal D} {\cal J}_{31}
{\cal J}_{43} +
k^2 {\cal J}_{41}', \nonumber \\
\widetilde{{\cal J}_{53}} & = &
{\cal J}_{53} - k \bar{{\cal D}} {\cal J}_{23} + k {\cal H}_{\bar{2}} {\cal D} {\cal J}_{
43} + {\cal J}_{13} {\cal J}_{21} +
{\cal J}_{13} {\cal J}_{31} {\cal J}_{43} - \nonumber \\
& & {\cal J}_{43} {\cal J}_{52} - k \bar{{\cal D}} {\cal J}_{13}
{\cal J}_{41} -
k {\cal D} {\cal H}_{\bar{2}} {\cal J}_{43} + k^2 {\cal J}_{43}',
\end{eqnarray}
The unconstrained $N=2$ stress tensor is given by
\begin{eqnarray}
{\cal T}=\frac{1}{k} {\cal J}_{13} {\cal J}_{31}-
\frac{1}{k} {\cal J}_{52}+\bar{{\cal D}} {\cal H}_{1}-
{\cal D} {\cal H}_{\bar{2}}
\end{eqnarray}
with central charge $2 k$. All the supercurrents are superprimary with
respect to ${\cal T}$.
After rescaling
\begin{eqnarray}
{\cal J}_{14} \rightarrow \frac{1}{k} {\cal J}_{14}, \;\;
{\cal J}_{34} \rightarrow \frac{1}{k} {\cal J}_{34}, \;\;
{\cal J}_{51} \rightarrow \frac{1}{k} {\cal J}_{51}, \;\;
{\cal J}_{53} \rightarrow \frac{1}{k} {\cal J}_{53},
\end{eqnarray}
we can write down the remaining SOPEs in the following form
\vspace{2cm}
\begin{eqnarray}
{\cal H}_{1}(Z_{1}) {\cal H}_{\bar{2}}(Z_{2}) & = &
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \frac{k}{2}-
\frac{1}{z_{12}} k, \nonumber \\
{\cal H}_{1}(Z_{1}) {\cal J}_{13}(Z_{2}) & = & \frac{\bar{\theta}_{12}}{z_{12}}
{\cal J}_{13}, \nonumber \\
{\cal H}_{1}(Z_{1}) {\cal J}_{31}(Z_{2}) & = & -\frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{31}, \nonumber \\
{\cal H}_{\bar{2}}(Z_{1})
{\cal J}_{13} (Z_{2}) & = &
\frac{\theta_{12}}{z_{12}} {\cal J}_{13}, \nonumber \\
{\cal H}_{\bar{2}} (Z_{1}) {\cal J}_{31}(Z_{2}) & = & -\frac{\theta_{12}}{z_{12}}
{\cal J}_{31}, \nonumber \\
{\cal J}_{13} (Z_{1}) {\cal J}_{31}(Z_{2}) & = & -
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \frac{k}{2} +
\frac{1}{z_{12}} k-
\frac{\bar{\theta}_{12}}{z_{12}} {\cal H}_{\bar{2}}-\frac{\theta_{12}}{z_{12}} {\cal H}_{1}+ \nonumber \\
& & \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[ -\bar{{\cal D}} {\cal H}_{1}-\frac{1}{k}
{\cal H}_{1} {{\cal H}_{\bar{2}}} + \frac{1}{k}
{\cal J}_{13} {\cal J}_{31} \right],
\label{eq:u2}
\end{eqnarray}
\begin{eqnarray}
& & {\cal H}_{1} (Z_{1})
\left\{ \begin{array}{ccl}
{\cal J}_{14} (Z_{2}) & = & \frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{14}, \nonumber \\
{\cal J}_{51} (Z_{2}) & = & -\frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{51}, \nonumber \\
\end{array} \right. \nonumber \\
& & {\cal H}_{\bar{2}} (Z_{1})
\left\{\begin{array}{ccl}
{\cal J}_{34} (Z_{2}) & = &
-\frac{\theta_{12}}{z_{12}} {\cal J}_{34}, \nonumber \\
{\cal J}_{53} (Z_{2}) & = &
\frac{\theta_{12}}{z_{12}} {\cal J}_{53}, \nonumber \\
\end{array} \right. \nonumber \\
& & {\cal J}_{13} (Z_{1})
\left\{\begin{array}{ccl}
{\cal J}_{14} (Z_{2}) & = &
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} {\cal J}_{13} {\cal J}_{14}, \nonumber \\
{\cal J}_{34} (Z_{2}) & = &
-\frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{14} - \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} {\cal H}_{1}
{\cal J}_{14}, \nonumber \\
{\cal J}_{51} (Z_{2}) & = &
\frac{\bar{\theta}_{12}}{z_{12}} {\cal J}_{53} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} \left[ {\cal H}_{1}
{\cal J}_{53} +
{\cal J}_{13} {\cal J}_{51} \right], \nonumber \\
\end{array} \right. \nonumber \\
& & {\cal J}_{31} (Z_{1})
\left\{\begin{array}{ccl}
{\cal J}_{14} (Z_{2}) & = &
\frac{\theta_{12}}{z_{12}} {\cal J}_{34} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k}
{\cal J}_{34} {\cal H}_{\bar{2}}, \nonumber \\
{\cal J}_{34} (Z_{2}) & = &
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} {\cal J}_{31} {\cal J}_{34}, \nonumber \\
{\cal J}_{53} (Z_{2}) & = &
-\frac{\theta_{12}}{z_{12}} {\cal J}_{51} - \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} \left[
{\cal H}_{\bar{2}}
{\cal J}_{51}+
{\cal J}_{31} {\cal J}_{53} \right], \nonumber \\
\end{array} \right.
\end{eqnarray}
\begin{eqnarray}
{\cal J}_{14} (Z_{1}) {\cal J}_{34} (Z_{2}) & = &
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \frac{1}{k} {\cal J}_{14} {\cal J}_{34}, \nonumber \\
{\cal J}_{14} (Z_{1}) {\cal J}_{51} (Z_{2}) & = &
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^3} k+
\frac{1}{z_{12}^2} k - \frac{\bar{\theta}_{12}}{z_{12}^2} {\cal H}_{\bar{2}} \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[
-\frac{1}{2} {\cal T} - \frac{1}{2} {\cal D} {\cal H}_{\bar{2}} -
\frac{1}{2 k} {\cal H}_{1} {\cal H}_{\bar{2}} +
\frac{2}{k} {\cal J}_{13} {\cal J}_{31} \right] \nonumber \\
& & +
\frac{1}{z_{12}} \left[ {\cal T} + {\cal D} {\cal H}_{\bar{2}} + \frac{1}{k}
{\cal H}_{1}
{\cal H}_{\bar{2}} - \frac{2}{k} {\cal J}_{13}
{\cal J}_{31} \right] \nonumber \\
& & +
\frac{\bar{\theta}_{12}}{z_{12}} \left[ \bar{{\cal D}} {\cal T} - \frac{1}{k} {\cal H}_{\bar{2}} {\cal T} -
\frac{1}{k} {\cal H}_{\bar{2}} {\cal D} {\cal H}_{\bar{2}} -
\frac{1}{k^2} {\cal J}_{13}
{\cal H}_{\bar{2}} {\cal J}_{31} \right. \nonumber \\
& & \left.
+ \frac{1}{k} \bar{{\cal D}} {\cal H}_{1} {\cal H}_{\bar{2}} -
\frac{1}{k} \bar{{\cal D}} {\cal J}_{13}
{\bar{
{\cal J}_{13}}} - {\cal H}_{\bar{2}}' \right]
- \frac{\theta_{12}}{z_{12}} \frac{1}{k} {\cal D} \left( {\cal J}_{13}
{\cal J}_{31} \right) \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[
-\frac{1}{k^3} {\cal H}_{1} {\cal J}_{13}
{\cal H}_{\bar{2}}
{\cal J}_{31} -
\frac{1}{k^2} {\cal H}_{1} \bar{{\cal D}} {\cal J}_{13}
{\cal J}_{31} +
\frac{1}{k^2} {\cal J}_{13} {\cal J}_{31}
{\cal T} \right. \nonumber \\
& &
\left. + \frac{1}{k^2} {\cal J}_{13} {\cal D} {\cal H}_{\bar{2}}
{\cal J}_{31} +
\frac{1}{k} \left( {\cal J}_{13} {\cal J}_{31} \right)' +
\frac{1}{k} {\cal J}_{34} {\cal J}_{53} +
\frac{1}{k} \bar{{\cal D}} {\cal J}_{13} {\cal D} {\cal J}_{31}
\right], \nonumber \\
{\cal J}_{14} (Z_{1}) {\cal J}_{53} (Z_{2}) & = &
\frac{{\theta}_{12}}{z_{12}^2} {\cal J}_{13} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[ \frac{3}{2} \bar{{\cal D}}
{\cal J}_{13} -
\frac{1}{2 k} {\cal J}_{13} {\cal H}_{\bar{2}}
\right] \nonumber \\
& & + \frac{1}{z_{12}}
\left[- \bar{{\cal D}} {\cal J}_{13} + \frac{1}{k} {\cal J}_{13}
{\cal H}_{\bar{2}}
\right] + \frac{\bar{\theta}_{12}}{z_{12}} \frac{1}{k}
\bar{{\cal D}} {\cal J}_{13} {\cal H}_{\bar{2}} \nonumber \\
& & + \frac{\theta_{12}}{z_{12}} \left[
-\frac{1}{k^2} {\cal H}_{1} {\cal J}_{13}
{\cal H}_{\bar{2}} + \frac{1}{k}
{\cal J}_{13} {\cal T} -
\frac{1}{k} \bar{{\cal D}} {\cal H}_{1} {\cal J}_{13} +
{\cal J}_{13}' \right] \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[ \bar{{\cal D}} {\cal J}_{13}'
-\frac{1}{k} \bar{{\cal D}} {\cal H}_{1} \bar{{\cal D}} {\cal J}_{13} -
\frac{1}{k^2} \bar{{\cal D}} \left( {\cal H}_{1} {\cal J}_{13}
{\cal H}_{\bar{2}} \right) +
\frac{1}{k} \bar{{\cal D}} \left( {\cal J}_{13} {\cal T} \right) \right], \nonumber \\
{\cal J}_{34} (Z_{1}) {\cal J}_{51} (Z_{2}) & = &
-\frac{\bar{\theta}_{12}}{z_{12}^2} {\cal J}_{31} + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[ -
\frac{1}{2} {\cal D} {\cal J}_{31} +
\frac{3}{2 k} {\cal H}_{1} {\cal J}_{31} \right] \nonumber \\
& & +
\frac{1}{z_{12}} \left[ {\cal D} {\cal J}_{31} - \frac{1}{k} {\cal H}_{1}
{\cal J}_{31} \right]
+ \frac{\theta_{12}}{z_{12}} \frac{1}{k} {\cal H}_{1}
{\cal D} {\cal J}_{31} \nonumber \\
& & +
\frac{\bar{\theta}_{12}}{z_{12}} \left[- \frac{1}{k^2} {\cal H}_{1} {\cal H}_{\bar{2}}
{\cal J}_{31} -
\frac{1}{k} {\cal J}_{31} {\cal T} -
\frac{1}{k} {\cal D}
{\cal H}_{\bar{2}}
{\cal J}_{31} -
{\cal J}_{31}' \right] \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[ \frac{1}{k^2} {\cal H}_{1}
{\cal J}_{31} {\cal T} -
\frac{1}{k^2} {\cal H}_{1} \bar{{\cal D}} {\cal H}_{1}
{\cal J}_{31} +
\frac{1}{k^2} {\cal H}_{1} {\cal D} {\cal H}_{\bar{2}}
{\cal J}_{31} \right. \nonumber \\
& & \left.
+ \frac{1}{k} \left( {\cal H}_{1} {\cal J}_{31} \right)'
-
\frac{1}{k} {\cal J}_{34} {\cal J}_{51} +
\frac{1}{k} \bar{{\cal D}} {\cal H}_{1} {\cal D} {\cal J}_{31} \right], \nonumber \\
{\cal J}_{34} (Z_{1}) {\cal J}_{53} (Z_{2}) & = &
-\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^3} k +
\frac{1}{z_{12}^2} k + \frac{{\theta}_{12}}{z_{12}^2} {\cal H}_{1} \nonumber \\
& & + \frac{\theta_{12} \bar{\theta}_{12}}{z_{12}^2} \left[
-\frac{1}{2} {\cal T} +
\frac{3}{2} \bar{{\cal D}} {\cal H}_{1} -
\frac{1}{2 k} {\cal H}_{1} {\cal H}_{\bar{2}}
\right] \nonumber \\
& & +
\frac{1}{z_{12}} \left[ {\cal T} - \bar{{\cal D}} {\cal H}_{1} + \frac{1}{k}
{\cal H}_{1}
{\cal H}_{\bar{2}} \right] +
\frac{\bar{\theta}_{12}}{z_{12}} \left[ \bar{{\cal D}} {\cal T} +
\frac{1}{k} \bar{{\cal D}} {\cal H}_{1} {\cal H}_{\bar{2}}
\right] \nonumber \\
& & + \frac{\theta_{12}}{z_{12}} \left[ \frac{1}{k} {\cal H}_{1} {\cal T} -
\frac{1}{k} {\cal H}_{1} \bar{{\cal D}} {\cal H}_{1} +
{\cal H}_{1}' \right] \nonumber \\
& & +
\frac{\theta_{12} \bar{\theta}_{12}}{z_{12}} \left[ \bar{{\cal D}} {\cal H}_{1}' -
\frac{1}{k} {\cal H}_{1} \bar{{\cal D}} {\cal T} +
\frac{1}{k} \bar{{\cal D}} {\cal H}_{1} {\cal T} -
\frac{1}{k} \bar{{\cal D}} {\cal H}_{1} \bar{{\cal D}} {\cal H}_{1} \right].
\end{eqnarray}
Let us summarize the $N=2$ $u(3)$ SCA.
It contains unconstrained spin $1$ $N=2$ stress tensor
${\cal T}$, the spin $1/2$ chiral and anti-chiral supercurernts
${\cal H}_{1}$ and $
{\cal H}_{\bar{2}}$, the spin $1/2$ supercurrents ${\cal J}_{
13}$ and ${\cal J}_{31}$ subjected to the nonlinear chirality constraints,
the spin $1$ anti-chiral supercurrent ${\cal J}_{53}$
and the spin $1$ constrained supercurrents ${\cal J}_{14}, {\cal J}_{51},
{\cal J}_{34}$. All these supercurrents are bosonic (fermionic) for integer
(half-integer) spin.
The supercurrents ${\cal H}_{1},{\cal H}_{\bar{2}}, {\cal J}_{13}, {\cal J}_{31}$ possess
a closed set of SOPEs (see eqs. (\ref{eq:u2})) and form $N=2$ $u(2)=u(2|0)$
current subalgebra.
We would like to note that in \cite{Ra} an $N=1$ superfield
extension of $u(3)$ KB SCA has been found.
The field content of both $N=1$ $u(3)$ SCA of Ref. \cite{Ra} and our
superalgebra is the same (modulo different choices of the basis for the
constituent currents), but the novel point is that
we have succeeded in arranging the relevant currents into $N=2$
supermultiplets
(by putting them into properly constrained $N=2$ supercurrents)
and thereby revealed $N=2$ supersymmetry of this superlagebra which was
hidden in the formulation of Ref. \cite{Ra}.
Let us now consider a secondary Hamiltonian reduction of $N=2$ $u(3)$
SCA to $N=2$ $W_{3}^{(2)}$ SCA. It goes as follows.
With respect to the new stress tensor ${\cal T}_{\mbox{new}}$
\begin{eqnarray}
{\cal T}_{\mbox{new}}={\cal T} - 2 \bar{{\cal D}} {\cal H}_{1}- {\cal D}
{\cal H}_{\bar{2}}
\end{eqnarray}
the supercurrent ${\cal J}_{14}$ has zero spin and $u(1)$ charge,
while the spin and $u(1)$ charge of ${\cal J}_{13}$ are equal, respectively,
to $-1/2$ and $-1$. Thus we can impose two first-class constraints
\begin{eqnarray}
{\cal J}_{14}=1, \;\; {\cal J}_{13}=0.
\label{eq:u314}
\end{eqnarray}
Gauge fixing procedure for either constraints can be done as usual.
So we fix the gauge by
\begin{eqnarray}
{\cal H}_{1}=0, \;\; {\cal J}_{34}=0 \;.
\label{eq:u3h1}
\end{eqnarray}
Using (\ref{eq:u314}), (\ref{eq:u3h1}) we see that (\ref{eq:u3cons})
is reduced to (\ref{eq:w32con}).
The dimensions and $u(1)$ charges of the surviving
supercurrents ${\cal J}_{31}, {\cal H}_{\bar{2}},
{\cal J}_{51}, {\cal J}_{53}$ with respect to ${\cal T}_{\mbox{new}}$ coincide with
those in Table.5.
Let us come back to discussion of $N=2$ $u(3)$ SCA. A simple
inspection of its current content shows that there are four
spin $1/2$ currents in it besides the
set of $16$ currents with higher spins. Like in the case
of $N=2$ $u(2|1)$ SCA, they can be factored out by passing to
a new basis where they (anti)commute with the remainder of the currents.
After decoupling of these spin $1/2$ currents our $N=2$ $u(3)$ SCA
reproduces $u(3)$ KB SCA \cite{K,Ber}.
Let us remind the current content of $u(3)$ KB SCA.
It is generated by $16$ currents: Virasoro stress tensor $
T_{KB}$, six spin $3/2$ currents $G^{a}_{KB}$ and
${\bar{G}}_{a\;KB}$,
and nine spin $1$ currents forming the $u(3)$ affine current algebra,
namely, $u(1)$ current $H_{KB}$ and eight $su(3)$ currents
${J^{a}}_{b\;KB}$ with zero trace $ ( {J^{a}}_{a\;KB} =0 ) $.
Indices $a, b$ are running from $1$ to $3$ and
correspond to the fundamental $3$ and its conjugate $\bar{3}$
representations of $su(3)$ (for upper and lower positions, respectively).
Below we give the precise correspondence
between these $u(3)$ KB SCA currents and components of the original set of
$N=2$ $u(3)$ SCA supercurrents
\begin{eqnarray}
{J^{3}}_{2,KB}
& = & \left( \bar{{\cal D}} {\cal J}_{13}-\frac{1}{k} {\cal J}_{13}
{\cal H}_{\bar{2}} \right) \vert, \nonumber \\
-{J^{2}}_{3,KB}
& = & \left( {\cal D} {\cal J}_{31}-\frac{1}{k} {\cal H}_{1}
{\cal J}_{31} \right) \vert, \nonumber \\
-\frac{1}{3} H_{KB}-{J^{2}}_{2,KB} & = &
\left( \bar{{\cal D}} {\cal H}_{1}-\frac{1}{k} {\cal J}_{13}
{\cal J}_{31} \right) \vert, \nonumber \\
\frac{1}{3} H_{KB}-{J^{1}}_{1,KB}-{J^{2}}_{2,KB} & = &
\left( {\cal D} {\cal H}_{\bar{2}} -\frac{1}{k} {\cal J}_{13}
{\cal J}_{31} \right) \vert, \nonumber \\
-k {J^{1}}_{2,KB} & = & {\cal J}_{14} \vert, \nonumber \\
\frac{k}{\sqrt{2}} {\bar{G}}_{2,KB} & = &
\left( \bar{{\cal D}} {\cal J}_{14}+\frac{1}{k} {\cal H}_{\bar{2}}
{\cal J}_{14} \right) \vert, \nonumber \\
-k {J^{1}}_{3,KB} & = & {\cal J}_{34} \vert, \nonumber \\
\frac{k}{\sqrt{2}} {\bar{G}}_{3,KB} & = &
\left( \bar{{\cal D}} {\cal J}_{34} +\frac{1}{k} {\cal J}_{31}
{\cal J}_{14} \right) \vert, \nonumber \\
-k {J^{2}}_{1,KB} & = & {\cal J}_{51} \vert, \nonumber \\
\frac{k}{\sqrt{2}} G^{2}_{KB} & = &
{\cal D} {\cal J}_{51} \vert, \nonumber \\
-\frac{k}{3} {H}_{KB}-k {J^{1}}_{1,KB} & = &
\left( -k {\cal T}+{\cal J}_{13} {\cal J}_{31}+k \bar{{\cal D}}
{\cal H}_{1}-k {\cal D} {\cal H}_{\bar{2}}-
{\cal H}_{1} {\cal H}_{\bar{2}} \right) \vert, \nonumber \\
\frac{k}{\sqrt{2}} G^{1}_{KB} & = &
\left( -k {\cal D} {\cal T}-\frac{1}{k} {\cal H}_{1} {\cal J}_{13}
{\cal J}_{31}-{\cal J}_{13} {\cal D}
{\cal J}_{31}+{\cal H}_{1} {\cal D} {\cal H}_{\bar{2}}
\right) \vert, \nonumber \\
\frac{k}{\sqrt{2}} {\bar{G}}_{1,KB} & = &
\left( -k \bar{{\cal D}} {\cal T}+\bar{{\cal D}} {\cal J}_{13} {\cal J}_{31}-
\frac{1}{k} {\cal J}_{13} {\cal H}_{\bar{2}}
{\cal J}_{31}-\bar{{\cal D}} {\cal H}_{1} {\cal H}_{\bar{2}}
\right) \vert, \nonumber \\
T_{KB} & = &
\left( -\frac{1}{2k} \left[ k [ {\cal D}, \bar{{\cal D}} ] {\cal T}+ {\cal H}_{1}
{\cal H}_{\bar{2}}'-
{\cal H}_{1}' {\cal H}_{\bar{2}}+ {\cal J}_{13}'
{\cal J}_{31}-
{\cal J}_{13} {\cal J}_{31}' \right] \right) \vert, \nonumber \\
-k {J^{3}}_{1,KB} & = & {\cal J}_{53} \vert, \nonumber \\
\frac{k}{\sqrt{2}} {G^{3}}_{KB} & = &
\left( {\cal D} {\cal J}_{53}+\frac{1}{k} {\cal H}_{1}
{\cal J}_{53}+\frac{1}{k}
{\cal J}_{13} {\cal J}_{51} \right)
\vert.
\end{eqnarray}
The OPEs of these currents are a particular case of OPEs of
$u(m|n)$ SCA given in Appendix C, eqs. (\ref{eq:comp}),
with the following correspondence
\begin{eqnarray}
k=\kappa, \;\; T_{KB}=T, \;\; H_{KB}={J^{a}}_{a}, \nonumber \\
{J^{a}}_{b, KB}={J^{a}}_{b}-\frac{1}{3} \delta^{a}_{b} {J^{c}}_{c}, \;\;
G^{a}_{KB}=i G^{a}, \;\; \bar{G}_{a,KB}=i {\bar{G}}_{a},
\end{eqnarray}
and $m=3, n=0$.
It is worth to notice that $G^{1}_{KB}, \bar{G}_{1,KB}$ are
related to the two fermionic components of linear $N=2$
superconformal stress tensor, ${\cal T}$, through nonlinear transformations.
So, two of six supersymmetries of $u(3)$ KB SCA are linearized
by passing to $N=2$ $u(3)$ SCA(viz., by adding four spin $1/2$ fermionic
currents), but four of them remain nonlinear.
\section{\bf Conclusion and outlook}
\setcounter{equation}{0}
In this paper we constructed $N=2$ $sl(n|n-1)^{(1)}$ current superalgebras
and developed a general scheme of classical
hamiltonian reduction in $N=2$ superspace. We applied it to
$N=2$ extension of affine superalgebra $sl(3|2)^{(1)}$.
As the main result, we deduced some new extensions
of $N=2$ SCA, $N=2$ $u(2|1)$ and $N=2$ $u(3)$ SCAs.
Within our scheme, these two new algebras turn out to be more
fundamental than the previously explored
$N=2$ $W_{3}^{(2)},$ $N=2$ $ W_{3}$ SCAs in the sense that the
latter can be generated by secondary hamiltonian reductions from the former.
The following diagram depicts basic points of our reduction procedure.
\begin{picture}(500,200)(-200,-170)
\setlength{\unitlength}{0.3mm}
\put(0,0){$N=2$ $sl(3|2)^{(1)}$}
\put(40,-4){\vector(-1,-1){45}}
\put(40,-4){\vector(1,-1){45}}
\put(-85,-70) {$N=2$ $u(2|1)$ $SCA$}
\put(45,-70) {$N=2$ $u(3)$ $SCA$}
\put(0,-135) {$N=2$ $W_{3}^{(2)}$ $SCA$}
\put(-20,-78) {\vector(1,-1){40}}
\put(100,-78) {\vector(-1,-1){40}}
\put(0,-200) {$N=2$ $W_{3}$ $SCA$}
\put(40,-140) {\vector(0,-1){45}}
\end{picture}
There are several problems to be worked out and questions
which at present are open.
Quantizing $W$ algebras associated with arbitray
embeddings of $sl(2)$ into (super)algebras has been studied
in \cite{ST}. These results were extended to $N=1$ affine Lie
superalgebras in superspace formalism \cite{MR}.
It is interesting to see whether the quantization of our superconformal
algebras can be carried out in $N=2$ superfield formalism.
It would be also interesting to study how $N=2$ $W_{4}$
\cite{YW}, and $N=2$ extensions (yet to be constructed)
of some other reductions of $sl(4)$ could come out in the
framework of hamiltonian reduction applied to $N=2$
$sl(4|3)^{(1)}$ superalgebra.
There exist some other superalgebras
which have completely fermionic simple root
system and admit $osp(1|2)$ principal embedding: $osp(2n \pm 1|2n),
osp(2n|2n), osp(2n+2|2n)$ $n \geq 1$ and $D(2,1; \alpha)$
$ \alpha \neq0,-1$ \cite{LSS}.
It is natural to apply our general procedure to these
superalgebras and see whether they admit $N=2$
superfield extensions.
It is also rather straightforward to construct free superfield
realizations for $N=2$ $u(2|1)$ and $N=2$ $u(3)$ SCAs. An interesting
related problem is to understand how these latter algebras reappear
in the $N=2$ superfield Toda and WZNW setting \footnote{For $N=2$ $W_n$
this is discussed in \cite{DM}.}.
It is rather exciting task to extend the techniques developed here
to the $N=4$ case, and, as a first step, to regain ``small''
$N=4$ SCA within the hamiltonian reduction framework in a
manifestly supersymmetric $N=4$ superfield fashion.
\vspace{1cm}
{\Large \bf Acknowledgments}
\vspace{0.5cm}
We would like to thank M. Magro, E. Ragoucy, A. Semikhatov,
P. Sorba, F. Toppan and, especially, F. Delduc and
S. Krivonos for many useful
and clarifying discussions.
We are grateful to V. Ogievetsky for his interest in this work.
Two of us (E.I. \& A.S.) acknowledge a partial support from the
Russian Foundation of Fundamental Research, grant 93-02-03821,
and the International Science Foundation, grant M9T300.
\vspace{1cm}
{\Large \bf Appendix A: Notations for $sl(2|1)$ superalgebra}
\setcounter{equation}{0}
\defD.\arabic{equation}{A.\arabic{equation}}
\vspace{0.5cm}
The generators of $sl(2|1)$ superalgebra in the complex basis introduced
in Section $2$ for the fundamental representation are given by
\begin{eqnarray}
t_{1}=
\left( \begin{array}{ccc}
0 & 1 & 0 \\
0 & 0 & 0 \\
0 & 0 & 0 \\
\end{array} \right) \;,
t_{2}=
\left( \begin{array}{ccc}
0 & 0 & 0 \\
0 & 1 & 0 \\
0 & 0 & 1 \\
\end{array} \right) \;,
t_{3}=
\left( \begin{array}{ccc}
0 & 0 & 1 \\
0 & 0 & 0 \\
0 & 0 & 0 \\
\end{array} \right) \;,
t_{4}=
\left( \begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 1 \\
0 & 0 & 0 \\
\end{array} \right) \;, \nonumber \\
t_{\bar{1}}=
\left( \begin{array}{ccc}
0 & 0 & 0 \\
1 & 0 & 0 \\
0 & 0 & 0 \\
\end{array} \right) \;,
t_{\bar{2}}=
\left( \begin{array}{ccc}
1 & 0 & 0 \\
0 & 0 & 0 \\
0 & 0 & 1 \\
\end{array} \right) \;,
t_{\bar{3}}=
\left( \begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 0 \\
1 & 0 & 0 \\
\end{array} \right) \;,
t_{\bar{4}}=
\left( \begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 0 \\
0 & 1 & 0 \\
\end{array} \right),
\end{eqnarray}
where Cartan generators $t_{2}, t_{\bar{2}}$ together with $t_{1}, t_{
\bar{1}}$ form the bosonic subalgebra $sl(2) \oplus u(1)$, while the
generators
$t_{3}, t_{\bar{3}}, t_{4}, t_{\bar{4}}$ are fermionic roots.
In this basis the structure constants of $sl(2|1)$ are
\begin{eqnarray}
{f_{2,1}}^{1} & = & -1,\;
{f_{2,\bar{1}}}^{\bar{1}}=1,\;
{f_{2,3}}^{3}=-1,\;
{f_{2,\bar{3}}}^{\bar{3}}=1, \nonumber \\
{f_{\bar{2},1}}^{1} & = & 1,\;
{f_{\bar{2},\bar{1}}}^{\bar{1}}=-1,\;
{f_{\bar{2},\bar{4}}}^{\bar{4}}=1,\;
{f_{\bar{2},4}}^{4}=-1, \nonumber \\
{f_{1,\bar{1}}}^{2}& = & -1, \;
{f_{1,\bar{1}}}^{\bar{2}}=1,\;
{f_{1,\bar{3}}}^{\bar{4}}=-1,\;
{f_{1,4}}^{3}=1,\; \nonumber \\
{f_{\bar{1},\bar{4}}}^{\bar{3}} & = & -1,
{f_{\bar{1},3}}^{4}=-1,\;
{f_{\bar{1},3}}^{4}=1,\; \nonumber \\
{f_{3,\bar{3}}}^{1} & = & 1,\;
{f_{3,\bar{4}}}^{1}=1 ,\;
{f_{4,\bar{3}}}^{\bar{1}}=1, \;
{f_{4,\bar{4}}}^{2}=1,\;
\end{eqnarray}
and nonzero elements of Killing metric are given by
\begin{eqnarray}
g_{1 \bar{1}}=-g_{2 \bar{2}}=g_{3 \bar{3}}=g_{4 \bar{4}}=1.
\end{eqnarray}
The explicit relations between affine supercurrents
${\cal J}_{a}, {\cal J}_{\bar{a}}$
in this basis and the entries ${\cal J}_{mn}$ of the $sl(2|1)$ superlagebra
valued affine supercurrent
introduced in Section $3$ are as follows
\begin{eqnarray}
{\cal J}_{1}={\cal J}_{12}, \;\; {\cal J}_{2} \equiv {\cal H}_{1}={\cal J}_{11},
\;\; {\cal J}_{3}={\cal J}_{13}, \;\;
{\cal J}_{4}={\cal J}_{23}, \nonumber \\
{\cal J}_{\bar{1}}={\cal J}_{21}, \;\; {\cal J}_{\bar{2}} \equiv {\cal H}_{\bar{1}}={\cal J}_{22}, \;\;
{\cal J}_{\bar{3}}={\cal J}_{31}, \;\; {\cal J}_{\bar{4}}={\cal J}_{32}.
\end{eqnarray}
\vspace{1cm}
{\Large \bf Appendix B: Notations for $sl(3|2)$ superalgebra}
\setcounter{equation}{0}
\defD.\arabic{equation}{B.\arabic{equation}}
\vspace{0.5cm}
We choose four Cartan generators
$t_{1}, t_{\bar{1}}, t_{2}, t_{\bar{2}}$ of $sl(3|2)$
superalgebra in the fundamental representation in the following form
\begin{eqnarray}
t_{1}=
\left( \begin{array}{ccccc}
0 & 0 & 0 & 0 & 0 \\
0 & 1 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 1 & 0 \\
0 & 0 & 0 & 0 & 0 \\
\end{array} \right), \;\;
t_{\bar{1}}=
\left( \begin{array}{ccccc}
1 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 1 & 0 \\
0 & 0 & 0 & 0 & 0 \\
\end{array} \right), \nonumber \\
t_{2}=
\left( \begin{array}{ccccc}
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 1 \\
\end{array} \right), \;\;
t_{\bar{2}}=
\left( \begin{array}{ccccc}
0 & 0 & 0 & 0 & 0 \\
0 & 1 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 1 \\
\end{array} \right).
\label{eq:Cartan}
\end{eqnarray}
Each of the~ remaining $10$ unbarred~ generator
$t_{a}, a=3, 4, \dots , 12$ is represented~ by a $5 \times 5$
supermatrix~ with the only~ non-zero entry $1$ on the
intersection of $m$-th line and $n$-th row
$(m=1, 2, 3, 4, n > m)$. The barred generators $t_{\bar{a}}$ have their
nonzero entries $1$ in the bottom triangular part. Using the explicit
form of $sl(3|2)$ generators we can find their
(anti)commutators, taking into account their statistics (the generators
with non-zero entries inside the diagonal $3 \times 3$ and $2 \times 2$ blocks
are bosonic, all others are fermionic). In the complex basis, they
satisfy the following graded commutators
\begin{eqnarray}
[ t_{a}, t_{b} \}= {F_{a b}}^{c} t_{c}, \;\;
[ t_{\bar{a}}, t_{\bar{b}} \} = {F_{\bar{a} \bar{b}}}^{\bar{c}}
t_{\bar{c}}, \;\;
[ t_{a}, t_{\bar{b}} \} ={F_{a \bar{b}}}^{c} t_{c} +
{F_{a \bar{b}}}^{\bar{c}} t_{\bar{c}}
\end{eqnarray}
From this we can read off all the structure constants ${F_{AB}}^{C}$ which
are $1$ or $-1$ (remember that ${f_{AB}}^{C}=(-1)^{(d_{A}+1) d_{B}}
{F_{AB}}^{C}$). Killing metric $g_{a \bar{b}}$ is given by
$Str(t_{a} t_{\bar{b}})$ where we take usual convention for
supertrace.
Just as an example, we write down nonzero elements of
$g_{a \bar{b}}$ for the subset (\ref{eq:Cartan})
\begin{eqnarray}
-g_{1 \bar{1}}= g_{1 \bar{2}}= -g_{2 \bar{2}}=1.
\end{eqnarray}
\vspace{1cm}
{\Large \bf Appendix C: $u(m|n)$ SCAs \cite{DTH} in terms of currents}
\setcounter{equation}{0}
\defD.\arabic{equation}{C.\arabic{equation}}
\vspace{0.5cm}
This algebra includes Virasoro stress tensor $T$, $2n$ spin
$3/2$ bosonic currents, $G^{a}, {\bar{G}}_{a}$, $2 m$
spin $3/2$ fermionic currents, $G^{b}, {\bar{G}}_{b}, a=m+1, m+2, \dots, m+n,
b=1, 2, \dots,
m $, $(m^2+n^2)$ spin $1$ bosonic currents, ${J^{c}}_{d}, c,d=1,2, \dots,
m, {J^{e}}_{f}, e=m+1, m+2, \dots, m+n, f=n+1, n+2, \dots, m+n,$
and $2mn$ spin $1$ fermionic ones,
${J^{g}}_{h}, {J^{i}}_{j}, g=m+1,m+2, \dots, m+n,
h=1, 2, \dots, n, i=n+1, n+2, \dots, m+n, j=1, 2, \dots, m $.
The total set of $(m+n)^2$ spin $1$ currents forms $u(m|n)$
current algebra. Spin $3/2$ currents transform under fundamental
and conjugate representaions of $u(m|n)$, for upper and lower
positions of the indices, respectively.
These currents satisify the following OPEs
\begin{eqnarray}
T(z) T(w) & = &
\frac{1}{(z-w)^4} 3 \kappa +\frac{1}{(z-w)^2} 2 T+
\frac{1}{(z-w)} T', \nonumber \\
T(z) {J^{a}}_{b}(w) & = &
\frac{1}{(z-w)^2} {J^{a}}_{b}+
\frac{1}{(z-w)} {J^{a}}_{b}', \nonumber \\
T(z) {G}^{a}(w) & = &
\frac{1}{(z-w)^2} \frac{3}{2} {G}^{a}+
\frac{1}{(z-w)} {{G}^{a}}', \nonumber \\
T(z) {\bar{G}}_{a}(w) & = &
\frac{1}{(z-w)^2} \frac{3}{2} {\bar{G}}_{a}+
\frac{1}{(z-w)} {\bar{G}}_{a}', \nonumber \\
{J^{a}}_{b} (z) {J^{c}}_{d} (w) & = &
\frac{1}{(z-w)^2} \left[ \frac{1}{2-(m-n)} (-1) ^{(d_{a}+
1)(d_{b}+1)+(d_{c}+1)(d_{d}+1)}
\delta^{a}_{b} \delta^{c}_{d} + \right. \nonumber \\
& & \left. (-1)^{(d_{a}+d_{b}+d_{c}+1)(d_{d}+1)}
\delta^{a}_{d} \delta^{c}_{b} \right] \kappa
+
\frac{1}{(z-w)} \left[ (-1)^{(d_{a}+1)(d_{b}+d_{c})}
\delta^{c}_{b} {J^{a}}_{d} - \right. \nonumber \\
& & \left. (-1)^{(d_{a}+1)(d_{d}+1)+(d_{b}+1)(d_{c}+1)+
(d_{d}+1)(d_{b}+1)+(d_{d}+1)(d_{c}+1)} \delta^{a}_{d} {J^{c}}_{b}
\right], \nonumber \\
{J^{a}}_{b} (z) G^{c} (w) & = & \frac{1}{(z-w)} \delta^{c}_{b} G^{a},
\nonumber \\
{J^{a}}_{b} (z) {\bar{G}}_{c} (w) & = & -\frac{1}{(z-w)} \delta^{a}_{c}
(-1)^{(d_{b}+1)d_{c}} {\bar{G}}_{b}, \nonumber \\
G^{a} (z) {\bar{G}}_{b} (w) & = &-\frac{1}{(z-w)^3} (-1)^{(d_{a}+1)(
d_{b}+1)} \delta^{a}_{b} 4 \kappa + \nonumber \\
& & \frac{1
}{(z-w)^2} \left[2 (-1)^{(d_{a}+1)(d_{b}+1)} \delta^{a}_{b} {J^{c}}_{c}-
4 {J^{a}}_{b} \right] + \nonumber \\
& &
\frac{1}{(z-w)} \left[(-1)^{(d_{a}+1)(d_{b}+1)} \delta^{a}_{b} {J^{c}}_{c}'-
2 {J^{a}}_{b}'-
2 (-1)^{(d_{a}+1)(d_{b}+1)} \delta^{a}_{b} T - \right. \nonumber \\
& & \frac{1}{\kappa} (-1)^{(
d_{a}+1)(d_{b}+1)} \delta^{a}_{b} {J^{c}}_{c} {J^{d}}_{d}+
\frac{2}{\kappa} {J^{c}}_{c} {J^{a}}_{b} + \nonumber \\
& &
\left(
-\frac{1}{\kappa} \delta^{a}_{c} \delta^{d}_{b}+\frac{1}{2 \kappa}
(-1)^{(d_{a}+1)(d_{b}+1)} \delta^{a}_{b}
\delta^{c}_{d} \right) \times \nonumber \\
& & \left. \left( (-1)^{d_{e}+1} {J^{c}}_{e} {J^{e}}_{d}+
(-1)^{(d_{c}+1)(d_{d}+d_{e})+(d_{d}+1)(d_{e}+1)} {J^{e}}_{d} {J^{c}}_{e}
\right)\right],
\label{eq:comp}
\end{eqnarray}
\vspace{1cm}
{\Large \bf Appendix D: A Different Realization of $sl(n|n-1)$}
\setcounter{equation}{0}
\defD.\arabic{equation}{D.\arabic{equation}}
\vspace{0.5cm}
We can realize superalgebra $sl(n|n-1)$ in a different, though
equivalent way by $(2n-1) \times (2n-1)$ supermatrix whose entries
${\cal J}_{\tilde k \tilde l}$ are related
to those ${\cal J}_{kl}$ in the standard realization according to the
following rule \cite{BLNW}
\begin{eqnarray}
\widetilde{k} & = & 2k-1, \widetilde{l} =2l-1 \;\; \mbox{if} \;\;
1 \leq k, l \leq n \nonumber \\
\widetilde{k} & = & 2(k-n), \widetilde{l} =2(l-n) \;\; \mbox{if} \;\;
n < k, l \leq 2n-1 \;.
\end{eqnarray}
This parametrization corresponds to choosing the system of purely
fermionic simple roots in $sl(n|n-1)$.
It is very convenient when studying embeddings
of $sl(2|1)$ into $sl(n|n-1)$: the former is identified with proper
$3 \times 3$ blocks in the $sl(n|n-1)$ supermatrix ${}^{\ddagger\ddagger}$.
\footnotetext{We are grateful to
F. Delduc for explaining us the merits of this realization.}
Using this convention, the set of hamiltonian reduction constraints
we dealt with in the $sl(2|1)$ case can be rewritten in the following
suggestive way
\begin{eqnarray}
N=2 \;SCA: \;\; {{\cal J}_{mn}}^{\mbox{constr}}=
\left( \begin{array}{ccc}
\ast & 0 & 1 \\
\ast & \ast & \ast \\
\ast & 1 & \ast
\end{array} \right)
(\ref{eq:sl21constr})
\Longrightarrow
\left( \begin{array}{ccc}
\ast & 1 & 0 \\
\ast & \ast & 1 \\
\ast & \ast & \ast
\end{array} \right).
\label{eq:diasl21}
\end{eqnarray}
This picture shows that the constraints are concentrated in the upper
triangular part of the supercurrent matrix, and this is true as well
for the $sl(3|2)$ constraints except for (\ref{eq:conso5}). We first
present the matrices of constraints for the cases of $N=2$ $W_3$, $N=2$
$W_3^{(2)}$, $N=2$ $u(2|1)$ and $N=2$ $u(3)$ SCAs
\begin{eqnarray}
N=2\; W_3: \;\;\;{\cal J}_{mn}^{\mbox{constr}}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & 0 \\
\ast & \ast & 0 & \ast & 1 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & 1 & \ast & \ast
\end{array} \right)
(\ref{eq:w3cons})
\Longrightarrow
\left( \begin{array}{ccccc}
\ast & 1 & 0 & 0 & 0 \\
\ast & \ast & 1 & 0 & 0 \\
\ast & \ast & \ast & 1 & 0 \\
\ast & \ast & \ast & \ast & 1 \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right) \label{w3}
\end{eqnarray}
\begin{eqnarray}
N=2\; W_3^{(2)}: \;\;{\cal J}_{mn}^{\mbox{constr}}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & 0 \\
\ast & \ast & 0 & \ast & 1 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right)
(\ref{eq:consw32})
\Longrightarrow
\left( \begin{array}{ccccc}
\ast & 1 & 0 & 0 & 0 \\
\ast & \ast & 1 & 0 & 0 \\
\ast & \ast & \ast & 1 & 0 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right) \label{w32}
\end{eqnarray}
\begin{eqnarray}
N=2\;u(2|1):\;\;{\cal J}_{mn}^{\mbox{constr}}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & 0 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right)
(\ref{eq:consu2})
\Longrightarrow
\left( \begin{array}{ccccc}
\ast & 1 & 0 & 0 & 0 \\
\ast & \ast & 1 & 0 & 0 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right)
\label{eq:diau21}
\end{eqnarray}
\begin{eqnarray}
N=2\;u(3):\;\;{\cal J}_{mn}^{\mbox{constr}}=
\left( \begin{array}{ccccc}
\ast & 0 & \ast & \ast & 0 \\
\ast & \ast & 0 & \ast & 1 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right)
(\ref{eq:cou3})
\Longrightarrow
\left( \begin{array}{ccccc}
\ast & \ast & 0 & 0 & \ast \\
\ast & \ast & 1 & 0 & 0 \\
\ast & \ast & \ast & 1 & 0 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right) \;.
\label{eq:diau3}
\end{eqnarray}
The supermatrices of constraints for two ``noncanonical'' cases
described in Subsection 5.3, respectively with
the constrained $N=2$ stress tensor and/or spin $0$
supercurrents present, are given by
\begin{eqnarray}
{\cal J}_{mn}^{\mbox{constr}}=
\left( \begin{array}{ccccc}
\ast & \ast & \ast & 1 & \ast \\
\ast & \ast & \ast & 0 & \ast \\
\ast & \ast & \ast & 0 & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & 1 & 0 & \ast
\end{array} \right)
(\ref{eq:conso5})
\Longrightarrow
\left( \begin{array}{ccccc}
\ast & 1 & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 0 & \ast & \ast & \ast \\
\ast & 0 & \ast & \ast & 1 \\
\ast & 0 & \ast & \ast & \ast
\end{array} \right) \label{1uncan}
\end{eqnarray}
\begin{eqnarray}
{\cal J}_{mn}^{\mbox{constr}}=
\left( \begin{array}{ccccc}
\ast & 0 & 0 & 1 & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & 1 & 0 & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right)
(\ref{eq:consn5})
\Longrightarrow
\left( \begin{array}{ccccc}
\ast & 1 & 0 & \ast & 0 \\
\ast & \ast & 1 & \ast & 0 \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast \\
\ast & \ast & \ast & \ast & \ast
\end{array} \right)\;. \label{2uncan}
\end{eqnarray}
These pictures clearly demonstrate the relations between different reductons
in accord with the diagram of Section 6. Also it is seen from them
that it is natural to treat all the considered cases in the language
of $sl(2|1)$ embeddings. The case of $N=2$ $W_3$ corresponds to the
principal embedding of $sl(2|1)$ into $sl(3|2)$ while $N=2$ $u(2|1)$ and
$N=2$ $u(3)$ ones to two inequivalent non-principal embeddings. It would be
interesting to understand from an analogous point of view the cases \p{w32},
\p{1uncan}, \p{2uncan}. It seems that in this way one could
explain some peculiar features of them (lacking of the superprimary basis
in the $N=2$ $W_3^{(2)}$ case, the presence
of constrained $N=2$ stress tensor and/or spin $0$ supercurrents in two
remaining cases). Note that the complete classification of
$sl(2|1)$ embeddings, at the component level and in the string theory
context, is undertaken in \cite{RSS}.
|
\section{Introduction}
At least 50$\%$ of T Tauri stars (TTs) appear to be surrounded by
circumstellar dust disks (\markcite{Strom et al.\ 1989};
\markcite{Beckwith et al.\ 1990}, hereafter BSCG;
\markcite{Andr\'e \& Montmerle 1994}; \markcite{Henning \& Thamm 1994};
\markcite{Osterloh \& Beckwith 1995}).
Global disk properties can be inferred from models of spectral
energy distributions (SED's) from infrared to millimeter wavelengths
(\markcite{Adams, Lada, \& Shu 1987};
\markcite{Beckwith \& Sargent 1993}; \markcite{Mannings \& Emerson 1994}).
Masses and sizes are
similar to those assumed for the
early solar nebula, suggesting that the disks may be protoplanetary
(cf.\ BSCG; \markcite{Beckwith \& Sargent 1993}).
However, the SED models rely on assumptions about disk morphology
and radial structure, and about the nature of the
constituent dust grains.
Grain size and composition in these potentially
planet-forming disks can be inferred from
the spectral index, $\beta$, of the dust opacity
(cf.\ \markcite{Pollack et al.\ 1994}) if
thermal radiation from grains
in the disk is optically thin. Sub-arcsecond
resolution is necessary to image disks directly and
measure properties on spatial scales of
$\simless$100 AU at the distance of the nearest star-forming regions.
At wavelengths longer than 3mm, the emission
is well into the Rayleigh-Jeans part of the Planck curve and
very likely to be optically thin.
Dust continuum radiation has been detected
from a number of TTs at $\lambda$ = 3~mm
(\markcite{Sargent \& Beckwith 1993} and references therein),
but no thermal emission has been
detected unambiguously at longer wavelengths (Mundy et al.\ 1993).
The required spatial resolution and mJy sensitivities
can now be achieved using the VLA at wavelengths of
7 mm.
DO Tauri is a young TTs in the Taurus star formation complex
at a distance of 140 pc (\markcite{Elias 1978};
\markcite{Kenyon, Dobrzycka, \&
Hartmann 1994}). Estimates of its age and mass
range from 1.6 to 6.0 $\times 10^5$ yrs
and 0.3 to 0.7 $M_{\sun}$ (BSCG;
\markcite{Hartigan, Edwards \& Ghandour 1995}), depending
on the theoretical tracks used to place it on the H-R diagram.
The spectral energy distribution is consistent with the presence of a
$\sim$0.01 $M_{\sun}$ circumstellar disk
(BSCG; \markcite{Beckwith \& Sargent 1991}, hereafter BS;
\markcite{Mannings \& Emerson 1994}).
Asymmetric, blue-shifted, [OI] and [SII]
forbidden line emission
(\markcite{Appenzeller, Jankovics \& \"Ostreicher 1984};
\markcite{Edwards et al.\ 1987}; \markcite{Edwards, Ray \& Mundt 1993})
is resolved as an optical jet at PA 70$^\circ$
(\markcite{Hirth et al.\ 1994}).
The jet is approximately orthogonal to the direction
of linear optical polarization, PA $\sim 170^\circ$
(\markcite{Bastien 1982}), and to
the long axis of CO (2$\to$1) emission
detected in aperture synthesis
images of DO Tau, PA $\sim 160^\circ$ (\markcite{Koerner \& Sargent 1995}).
Kinematic models of the molecular line emission are consistent
with the presence of a circumstellar disk that is
centrifugally supported within a radius of 350 AU from
DO Tau (\markcite{Koerner 1994}).
Here, we report on sub-arcsecond images of the $\lambda$ = 7 mm
emission from DO Tau which were made using the recently
upgraded Very Large Array (VLA) of the National Radio Astronomy
Observatory\footnote{NRAO is operated by
Associated Universities Inc.\ under cooperative
agreement with the National Science Foundation.}.
We have supplemented these measurements with continuum
observations of DO Tau at other wavelengths to sample the spectral
distribution of emission from $\lambda$ = 1.3~mm to 3.6~cm
and improve our understanding of grain properties
in the circumstellar material.
\section{Observations and Results}
The VLA was used to observe DO Tau in radio
continuum emission at 43.3, 22.5, and 8.4 GHz
($\lambda$ = 7 mm, 1.3, and 3.6 cm).
The phase center was offset 1$''$ in both RA and Dec
from the stellar position of DO Tau
(\markcite{Herbig \& Bell 1988}), and the
total bandwidth was 100 MHz in right and left circular
polarizations. As for all observations discussed below,
molecular line emission is negligible within the narrow
band. Observations at 43.3 GHz were carried out on 1994
April 3--4 with the inner seven antennas
of the high-resolution A~configuration,
and on 1994 August 20 with 10 inner antennas of the B configuration.
Baselines up to 5.6 km provided UV coverage in the range
30--800 k$\lambda$.
On both dates, DO Tau was observed at 22.5 and 8.4 GHz
using the remainder of the VLA's
27 antennas. UV coverage was 50--2700 k$\lambda$ at
22.5 GHz and 20--1000k$\lambda$ at 8.4 GHz.
Absolute flux densities were calibrated
using 3C48 and 3C286 with an estimated uncertainty
of 20$\%$. At 43.3 GHz, gain calibration was
accomplished with periodic observations of 0333+321
with a measured flux density of 0.98 $\pm$ 0.06 Jy.
At 22.5 GHz and 8.4 GHz, the gain calibrator was
0400+258 with flux densities 0.65 $\pm$ 0.03 Jy
and 0.83 $\pm$ 0.01 Jy, respectively.
Data calibration and mapping used standard
routines in the NRAO AIPS software package.
Daytime atmospheric phase fluctuations during
A array observations necessitated
extensive editing and application of
a Gaussian taper to the UV data, resulting
in a $0.68'' \times 0.53''$ (FWHM)
synthesized beam at PA $-78^\circ$.
This corresponds to 95$\times$74 AU at DO Tau.
Fig.\ 1a displays the CLEANed image
of DO Tau at 43.3 GHz. An unresolved source
with flux density 1.80 $\pm 0.71$ mJy is detected at the
stellar position, $\alpha$(1950) = $04^h35^m24.19^s$,
$\delta$(1950) = $26^\circ 04' 54.5''$.
The $\pm 0.71$ mJy uncertainty includes
rms variations in the map ($\pm$ 0.35 mJy bm$^{-1}$)
and a possible 20\% error in absolute flux calibration.
At 22.5 and 8.6 GHz, DO Tau was not detected
within the area encompassed by the 43.3 GHz
synthesized beam to 3$\sigma$ levels of 0.76 and 0.17 mJy, respectively.
Observations were made with the Owens Valley millimeter array
at 89.2, 111.2, 221.5, and 232.0 GHz
(corresponding to $\lambda$ = 3.4, 2.7, 1.4, and 1.3 mm)
between 1993 September
and 1995 March. Measurements at 89 GHz were made with six
telescopes; four were used at 110 GHz, and five at
220 and 230 GHz. Overall UV-ranges were 5--60 k$\lambda$ (89 GHz),
5--25 k$\lambda$ (110 GHz), and
10--55 k$\lambda$ (220 Hz and 230 GHz).
Resulting FWHM synthesized beams are listed in Table I.
Antenna gains were determined from periodic
observations of 0528+134 and absolute flux density
calibration was based on measurements of Uranus.
Data were calibrated
using the Owens Valley software package, MMA, and mapped with
AIPS. At all four frequencies,
continuum emission is unresolved
and peaks at the position of the VLA 43.3 GHz image.
Aperture synthesis maps at 89 and 220 GHz are displayed
in Fig.\ 1b and 1c. Measured flux densities at all frequencies
are listed in Table I and displayed in Fig.\ 2.
\section{ Modeling and Discussion}
Our measurements of DO Tau between 8.4 and 230
GHz can be fit by a single power law,
F$_\nu$ $\propto \nu^{\alpha}$,
with index $\alpha$ = 2.39$\pm$0.23,
shown as a solid line in Fig.\ 2.
Earlier detections of radio emission
from TTs at wavelengths greater than 1.3 cm yielded
values of $\alpha$ between 0 and 1
(\markcite{Bieging, Cohen \& Schwartz 1984}) and have been
attributed to free-free
emission from ionized outflows (\markcite{Reynolds 1986}).
Extrapolating from the upper limit of
8.4 GHz emission with $\alpha$ = 1 (dotted line in figure 2),
we estimate that no more than 49$\%$ of DO Tau's 43.3 GHz emission
can arise as free-free radiation from an ionized jet.
Fig.\ 2 suggests the observed mm-wave flux
originates entirely from circumstellar dust.
The frequency dependence
of the mm-wave dust opacity, $\beta$,
can be derived from $\alpha$, since
$\alpha \approx 2 + \beta/(1 + \Delta)$, where $\Delta$
is the ratio of optically thick to optically thin emission from the
disk (BS, eqn.\ 1).
At frequencies where emission from a circumstellar disk
is largely optically thin and the Rayleigh-Jeans approximation
holds, $\beta \approx \alpha - 2$.
For DO Tau, continuum emission appears to be largely
optically thin, even at sub-millimeter wavelengths
(BS; \markcite{Mannings \& Emerson 1994}).
We estimate $0.39 \pm 0.23$ for $\beta_{1-7mm}$, in good agreement
with the BS value of $\beta_{0.6-1mm}$, $0.4 \pm 0.2$.
There is no evidence for the change in $\beta$ longward
of 2mm, postulated by \markcite{Mundy et al.\ (1993)}
for a few other TTs.
An estimate of $\beta$ can also be obtained by fitting the spectral
distribution of luminosity $L_\nu = 4 \pi D^2 \nu F_\nu$ with
a disk model which takes into account any contribution
from optically thick emission.
Following \markcite{BSCG} and \markcite{Adams et al.\ (1990)}, we assumed
power-law radial profiles in disk temperature and surface
density, T = T$_0 (R/R_0)^{-q}$ and $\Sigma = \Sigma_0 (R/R_0)^{-p}$
with p = 1.5 or 1.75.
The millimeter-wave emissivity of the grains, $\kappa$,
is 0.1$(\nu/10^{12} Hz)^\beta$ cm$^2$ g$^{-1}$.
The outer radius, R$_d$,
was allowed to take on values between 22 and 350 AU;
the former is the lower limit to disk size if all 1 mm
emission is optically thick (BS); the latter is the deconvolved half-maximum
radius of the CO-emitting region from aperture synthesis images.
These suggest a disk
inclination angle, $\theta$, of 40$^\circ$
(\markcite{Koerner \& Sargent 1995}).
{}From the 12, 25, and 60 $\mu$m IRAS fluxes, which
probe optically thick regions of the disk, we obtain
T = 218 K at 1 AU with q = 0.54, very close to the value derived
by BSCG for a face-on disk. Best-fit values of
$\beta$ and M$_d$, the total disk mass, were estimated
from the minimum reduced $\chi^2$ value.
Acceptable fits, with $\chi^2$ falling within
$\Delta\chi^2$ = 1 of its minimum value, were found for our entire
range of p and R$_d$ values. The best-fit model,
with $\beta = 0.6 \pm 0.3$, ${\rm M}_d = 1.0\pm0.5 \times 10^{-2}$
$M_{\sun}$, p = 1.75, R$_d$ = 350 AU, and
$\chi^2$ = 0.77, is plotted in Fig.\ 3 as a solid line,
along with the luminosity distribution derived from IRAS,
sub-millimeter, and millimeter observations of DO Tau.
Following BSCG (eqn.\ 20), these parameters yield
$\Delta \approx 0.28$ at $\lambda$ = 3 mm
and make possible a revised estimate of
$\beta$ from the power-law fit to data presented here.
For $\Delta = 0.28$ and $\alpha = 2.39$, we obtain
$\beta = 0.50 \pm 0.23$, in good agreement with the value obtained
from both our disk-model fit and that of
\markcite{Mannings \& Emerson (1994)}.
For the ISM, it is commonly assumed that $\beta$ is about 2 in the
millimeter wavelength regime (\markcite{Mathis 1990}).
However, a value of 1.3 has been obtained in recent
laboratory studies (\markcite{Agladze et al. 1994}). Even
lower values are suggested by sub-millimeter observations of T Tauri
disks (BS; \markcite{Mannings \& Emerson 1994}). A variety of explanations have
been proposed, including chemical composition,
physical shape, and grain growth
(\markcite{Wright 1987}; BS; \markcite{Kr\"ugel \&
Siebenmorgen 1994}; \markcite{Ossenkopf \& Henning 1994};
\markcite{Pollack et al.\ 1994}).
For DO Tau, we find $\beta \approx$ 0.5 and contend that our denser
sampling of the sub-millimeter regime and spectral coverage extending
to longer wavelengths effectively eliminates uncertainties that may
have complicated other derivations.
Grain properties in circumstellar disks are unlikely to display
the exotic range of chemical composition and physical shapes
required to reproduce this result in the laboratory.
By contrast, the growth of grain size distributions to
include particles larger than 1 mm accounts for our value of $\beta$
(cf.\ \markcite{Miyake and Nakagawa 1993}) and is
consistent with the short theoretical timescales ($\sim 100$ yr)
for production of mm-size particles in the early solar nebula
(cf. Fig.\ 19, \markcite{Cuzzi, Dobrovolskis, \& Champney 1993}).
If the average grain size in disks steadily increases due to
planetesimal formation, $\beta$ should decrease monotonically with age.
However, DO Tau is relatively young, only a
few $\times$ 10$^5$ yrs, with an outflow typical of an active
disk. Grain growth appears to have already occurred by the
early T-Tauri phase. Recent 7 mm images of a very young protostar,
HH24MMS (\markcite{Chandler et al.\ 1995}), also yield a lower value
of $\beta$ than found for many older TTs in sub-millimeter surveys
(cf.\ BS). These results are inconsistent with a simple
picture of gradually decreasing $\beta$; they could be explained
if mm-size grains grow quickly, followed by
generation of a new population of small dust grains by
planetesimal collisions (\markcite{Lissauer \& Stewart 1993}).
Long-wavelength observations of a statistical sample of
TTs disks encompassing a range of ages are clearly required
to test this hypothesis.
\acknowledgments
We are grateful to D. Wood for assistance during the first season of
43 GHz observations at the VLA.\ \ D.W.K.\ acknowledges support for
this work from NASA grant NGT-51071.
The Owens Valley millimeter-wave array is supported by NSF
grant AST-9314079. Research by A.I.S.\ on protoplanetary disks
is furthered by NASA grant NAGW-4030 from the ``Origins of
Solar Systems'' program. The research described in this
paper was carried out by the Jet Propulsion Laboratory,
California Institute of Technology,
and was sponsored by a fellowship from the National Research
Council and the National Aeronautics and Space Administration.
\begin{table}
\begin{center}
\begin{tabular}{ccccccc}
\tableline
\tableline
Frequency & Flux Density & Statistical & Total &
\hfil Synthesized &
Beam & Parameters \hfil \\
(GHz) & (mJy) & error (mJy) & error (mJy) & B$_{maj}$
& B$_{min}$ & PA \\
\tableline
\tableline
\phantom{00}8.4 & $<$0.17 (3$\sigma$) & ... & ... & 0.41$''$ & 0.38$''$
& 120$^\circ$ \\
\phantom{0}22.5 & $<$0.76 (3$\sigma$) & ... & ... & 0.25$''$ & 0.23$''$
& $-10^\circ$\\
\phantom{0}43.3 & \phantom{00}1.80 & 0.35 & 0.71 & 0.68$''$ & 0.53$''$
& $-78^\circ$ \\
\phantom{0}89.2 & \phantom{0}14.2\phantom{0} & 0.9 & 3.74 & 2.86$''$
& 2.10$''$ & 82$^\circ$\\
111.2 & \phantom{0}30.2\phantom{0} & 4.5 & 10.5 & 13.0$''$ & 5.4$''$
& $-69^\circ$\\
221.5 & \phantom{0}98.6\phantom{0} & 4.9 & 24.2 & 3.98$''$ & 3.15$''$
& 62$^\circ$\\
232.0 & 137.5\phantom{0} & 4.9 & 32.4 & 3.41$''$ & 3.15$''$ & $-84^\circ$\\
\tableline
\end{tabular}
\end{center}
\bigskip
\tablenum{1}
\caption{Radio (VLA) and mm-wave (OVRO) Continuum Flux Densities from
DO Tauri. Total errors include 20\% uncertainty in the absolute
flux density calibration.}
\end{table}
\clearpage
|
\section{Introduction}
In a previous investigation \cite{KW94}, we studied the N-N
interaction in the
framework of the chromo-dielectric soliton model from a static point
of view: we used the Born-Oppenheimer approximation to derive an
adiabatic N-N potential, which showed a soft core repulsion due
essentially to the color-electrostatic part of the one-gluon exchange.
Previous studies of the N-N interaction in terms of quark degrees of
freedom \cite{Fa83} have pointed out the importance of dynamical
methods (such as Generator Coordinate or Resonating Group) in the
calculation of a realistic N-N potential. For example, in a preceding
application of the non-topological soliton model to the N-N problem,
Schuh et al. \cite{Sc86} showed that a significant part of the
repulsion was due to dynamics; the absence of a repulsive core in some
previous works was also interpreted as an artifact of the adiabatic
approximation \cite{Fa83}.
The Lagrangian of the chromo-dielectric model is defined as in Ref.
\cite{KW94}:
\begin{equation}
{\cal L}={\cal L}_q+{\cal L}_\sigma+{\cal L}_G \ ,
\label{lag}
\end {equation}
with
\begin{eqnarray}
{\cal L}_q & = & \bar{\psi}\left(i\gamma^\mu D_\mu
- m_q\right)\psi \ , \\
{\cal L}_\sigma & = & \frac{1}{2}\partial{_\mu}\sigma\partial{^\mu}
\sigma-U\left(\sigma\right) \ , \\
{\cal L}_G & = & -\frac{1}{4} \kappa(\sigma) F^a_{\mu \nu}
F^{\mu\nu}_a \ ,
\end{eqnarray}
where $\psi$ is the quark operator and $m_q$ the current quark mass
matrix, set here to $m_q = 0$. The quark Lagrangian ${\cal L}_q$ is
expressed in terms of the covariant derivative
$D_\mu=\partial_\mu -ig_s T^a A_\mu^a$, and
$F^a_{\mu \nu}=\partial_\mu A_\nu^a-\partial_\nu A_\mu^a +
g_s f^{abc} A_\mu^b A_\nu^c$ is the $SU(3)$-color
tensor, where $f^{abc}$ are the $SU(3)$ structure constants and $T^a$
the $SU(3)$ generators. The quantity $U(\sigma)$ is the
self-interaction of the scalar field, $\sigma$, taken to be of
the form:
\begin{equation}
U(\sigma) = \frac{a}{2!}\sigma^2+\frac{b}{3!}\sigma^3+\frac{c}{4!}
\sigma^4+B \ ,
\end{equation}
and the dielectric function $\kappa(\sigma)$ is:
\begin{equation}
\kappa(\sigma)= 1+\theta(\sigma)\left(\frac{\sigma}{\sigma_v}
\right)^{\!2}\left[2\frac{\sigma}{\sigma_v}-3\right] \ ,
\end{equation}
where $\sigma_v$ is the scalar field's vacuum expectation value and
$\theta$ the usual step function.
The quark self-energy, due to interactions with confined gluons in the
dielectric medium, generates an effective coupling
between the quarks and the scalar field:
\begin{equation}
{\cal L}_{q\sigma} = -g_{eff}(\sigma)\bar{\psi}\psi \ ;
\end{equation}
we choose $g_{eff}(\sigma)$ to be of the form:
\begin{equation}
g_{eff}(\sigma)=g_0\sigma_v\left(\frac{1}{\kappa(\sigma)}-1\right) \ .
\label{cou}
\end{equation}
The expression in Eq. (\ref{cou}) is an approximation to what has been
calculated in Ref. \cite{Kr88}, and it is constructed to simulate
spatial confinement already at the mean field level. Note that the
coupling in Eq. (\ref{cou}) breaks the chiral invariance of the
Lagrangian of Eq. (\ref{lag}). This is an example of dynamical
symmetry breaking from which a massless Goldstone boson emerges
naturally.
The parameters involved in our calculation are $a,\ b,\ c,\ g_0$ and
$\alpha_s=g_s^2/4\pi$,
as discussed in detail in Ref. \cite{KW94}. By fitting the
nucleon and the $\Delta$ masses and the proton's rms charge radius
one remains with two free parameters, for which it is convenient to
use the dimensionless quantities c and $f=b^2/ac$. In this paper, we
have chosen the set $f=\infty$ and c=10000 taken from Table 1 of Ref.
\cite{KW94}. Contrary to Ref. \cite{Sc86}, the quarks here are not
only coupled to the $\sigma$-field but also interact among themselves
through one-gluon exchange (OGE). The OGE is treated in Abelian
approximation, and it can be separated into two terms: a
self-interaction term (in addition to $g_{eff}(\sigma)$ of Eq.
(\ref{cou})), which is required for color confinement and which
contributes to the one-body part of the Hamiltonian, and a term of
mutual interactions, which gives rise to the two-body part of the
Hamiltonian. As mentioned earlier, in the adiabatic approximation of
Ref. \cite{KW94}, it was the color-electrostatic part of the OGE,
which arises from the time-component of the gluonic quadrivector
$A_\mu^c$, and especially the corresponding self-energy diagrams,
which were responsible for the soft-core repulsion.
In this work, we incorporate the dynamics of the N-N interaction by
employing the Generator Coordinate Method (GCM); we derive an
approximate differential equation for the N-N wave function describing
the relative motion of the two nucleons.
This equation contains a local N-N
potential and an effective, coordinate dependent mass. By means of a
Fujiwara transformation, we then define a N-N separation length, x,
from the deformation parameter used previously in the adiabatic
approximation. This allows us to introduce a constant mass and to
rewrite the effective potential in terms of this coordinate x. One of
our objectives is to study the explicit role of the one-gluon exchange
effects on the local N-N potential, included for the first time in
such type of calculations. Another aim is to establish a connection
between our effective deformation parameter and the true internucleon
separation. The latter will enable us to apply our six-quark wave
functions to studies of the quark substructure of light nuclei, as
has been carried out already, for instance, in Ref. \cite{KW95}. The
present numerical results correspond to the (TS)=(10) sector, although
the formalism at hand can easily be extended to other isospin-spin
channels.
\section{The Generator Coordinate Method}
The GCM was introduced in the fifties by Hill and Wheeler \cite{HW53}
to describe collective motion in nuclear systems, such as rotation,
vibration or center of mass motion \cite{GW57,PY57}. Starting from a
many-body wave function $|\,\alpha\,\rangle$ depending on a collective
coordinate $\alpha$ (the deformation parameter of the six-quark system
in our case), a trial wave function is constructed by taking a linear
combination of the states $|\,\alpha\,\rangle$ with some weight
function $\Phi(\alpha)$,
\begin{equation}
|\,\Psi\,\rangle=\int\Phi(\alpha)~|\,\alpha\,\rangle~d\alpha \ ,
\end{equation}
where $\Phi(\alpha)$ is determined through the variational principle
\begin{equation}
\begin{array}{ccccc}
\delta E & = & \displaystyle \frac{\delta}{\delta\Phi^*}\frac{
\langle\,\Psi\,|\,H\,|\,\Psi\,\rangle}{\langle\,\Psi\,|\,\Psi\,
\rangle} & = & 0 \ , \\
\end{array}
\end{equation}
which leads to the Hill-Wheeler integral equation:
\begin{equation}
\int\langle\,\alpha\,|\,H-E\,|\,\alpha'\,\rangle~\Phi(\alpha')~
d\alpha' = 0 \ .
\end{equation}
This is a homogeneous Fredholm-type equation of the first kind,
notoriously unstable numerically. Although some methods exist to make
it stable (such as regularization \cite{WW88}, removal of the zero
normalization eigenmodes \cite{RS80}, Gaussian transform \cite{GT73},
etc.), we prefer to solve a differential equation approximately
equivalent to the Hill-Wheeler equation, both for numerical stability
and to facilitate comparison with analyses based on the Schr\"odinger
equation. In general, $\alpha$ is a multidimensional parameter. It is
at least three-dimensional when correspondence is made to ${\bf r}$.
We here restrict the calculations to the zero-impact parameter case,
which reduces the problem to a one-dimensional one, and leave
consideration of the angles to a later study.
\section{The Hill-Wheeler differential equation}
To derive such a differential equation, it is more convenient to work
with mean and relative deformation parameters, $\beta$ and $\delta$,
defined as
\begin{equation}
\begin{array}{ccc}
\beta & = & \displaystyle \frac{\alpha +\alpha'}{2} \ ,\\
\delta & = & \alpha - \alpha' \ .
\end{array}
\end{equation}
Expanding the weight function in a Taylor series around $\delta=0$,
one has:
\begin{eqnarray}
\langle\,\Psi\,|\,H-E\,|\,\Psi\,\rangle~= \int &&d\beta \int d\delta
\left[\Phi^*(\beta)+\frac{\delta}{2}{\Phi^*}'(\beta)+
\frac{\delta^2}{8}{\Phi^*}''(\beta)+\ldots\right]\nonumber\\
&&\langle\,\beta+\frac{\delta}{2}\,|\,H-E\,|\,\beta-\frac{\delta}
{2}\,\rangle\left[\Phi(\beta)-\frac{\delta}{2}\Phi'(\beta)+
\frac{\delta^2}{8}\Phi''(\beta)+\ldots\right] \ .
\end{eqnarray}
It is convenient to introduce the moments:
\begin{eqnarray}
H_n & = & \int d\delta~\langle\,\beta+\frac{\delta}{2}\,|\,H\,|\,
d\beta-\frac{\delta}{2}\,\rangle~\delta^n \ , \\
N_n & = & \int d\delta~\langle\,\beta+\frac{\delta}{2}\,|\,
d\beta-\frac{\delta}{2}\,\rangle~\delta^n \ .
\end{eqnarray}
Because $\langle\,\beta+\delta/2\,|\,H-E\,|\,\beta-\delta/2\,\rangle$
is an even function of $\delta$, the odd moments are zero. Supposing,
moreover, that $\langle\,\beta+\delta/2\,|\,H-E\,|\,\beta-\delta/2\,
\rangle$ is a sharply peaked function of $\delta$, one can stop the
expansion at second order in $\delta$. Partial integration and
variation by $\delta \Phi^*$ leads then to the Hill-Wheeler
differential equation:
\begin{equation}
\frac{1}{2}\frac{d}{d\beta}\left((H_2-EN_2)\frac{d\Phi}{d\beta}\right)
+ \left[H_0+\frac{1}{8}\frac{d^2H_2}{d\beta^2}\right]\Phi = E
\left[N_0+\frac{1}{8}\frac{d^2N_2}{d\beta^2}\right]\Phi \ .
\label{eqhw1}
\end{equation}
The introduction of a new function into the hermitian,
\begin{equation}
\tilde{\Phi}(\beta)=\sqrt{\tilde{N}_0(\beta)}~\Phi(\beta) \ ,
\end{equation}
where
\begin{equation}
\tilde{N}_0 = N_0 + \frac{1}{8}\frac{d^2 N_2}{d\beta^2} \ ,
\end{equation}
allows us to transform Eq. (\ref{eqhw1}) into hermitian form:
\begin{equation}
\left[-\frac{d}{d\beta}\frac{1}{2B(\beta)}\frac{d}{d\beta}+V(\beta)
\right]\tilde{\Phi}(\beta) = E\tilde{\Phi}(\beta) \ ,
\label{schrod}
\end{equation}
where $V(\beta)$ is given by:
\begin{equation}
V(\beta) = \frac{\tilde{H}_0}{\tilde{N}_0}+
\frac{1}{2\sqrt{\tilde{N}_0}}\frac{d}{d\beta}
\left((H_2 - EN_2)\frac{d}{d\beta}
\left(\frac{1}{\sqrt{\tilde{N}_0}}\right)\right) \ ,
\label{pot1}
\end{equation}
with
\begin{equation}
\tilde{H}_0 = H_0 + \frac{1}{8}\frac{d^2 H_2}{d\beta^2} \ .
\end{equation}
The term $B(\beta)$ is the effective mass:
\begin{equation}
B(\beta) = -\frac{\tilde{N}_0}{H_2 - EN_2} \ .
\label{mass}
\end{equation}
The total energy E enters the definition of B; its asymptotic form
at threshold is:
\begin{equation}
E=2m_N \ ,
\end{equation}
where $m_N$ is the nucleon mass.
Note that because we didn't incorporate
center of mass corrections the asymptotic value of the potential in
Eq. (\ref{pot1}) is not equal to the experimental value of $2 m_N$.
We have indeed $V(\infty)$ = 2468 MeV when gluons are not included and
$V(\infty)$ = 2240 MeV when they are. In practice, we could obtain a
value closer to the experimental value by subtracting recoil
corrections from the asymptotic energy:
\begin{equation}
m_{N}^{2}=\left(\frac{V(\infty)}{2}\right)^{\!2}
~-~\langle\,P^2\,\rangle \ ,
\end{equation}
but we prefer to avoid this step. This simplification does not affect
our conclusions. Following Brink and
Banerjee \cite{BB73}, we replace $E$ in the mass term by:
\begin{equation}
E \rightarrow \frac{H_0(\beta)}{N_0(\beta)} \ .
\label{repl}
\end{equation}
This approximation is consistent with neglecting higher order
derivatives of the moments in the Hamiltonian.
The moments $H_n$ (n=0, 2) and the corresponding quantities $B(\beta)$
and $V(\beta)$ have been calculated for three distinct cases:
\begin{equation}
\begin{array}{cccc}
(a)& H &=& H_{1}^{bag} + H_{OGE} \ , \\
(b)& H &=& H_{1}^{bag} + H_{OGE}^{mag} \ , \\
(c)& H &=& H_{1}^{bag} \ ,
\end{array}
\label{case}
\end{equation}
where $H_{1}^{bag}$, $H_{OGE}^{mag}$ and $H_{OGE}$ are, respectively,
the non-gluonic one-body term of the Hamiltonian, the color-magnetic
and the full one-gluon exchange contribution; they are given
explicitly in Ref. \cite{KW94}. In case (c), the one-gluon
exchange was left out
altogether. This is in the spirit of an earlier investigation where
the Friedberg-Lee soliton model was applied to N-N scattering
without considering gluonic degrees of freedom \cite{Sc86}.
In case (b), the
color-magnetic hyperfine interaction was accounted for, and in case
(a) the full color-magnetic and color-electrostatic OGE was included.
The reason to distinguish between cases (a) and (b) is that in the
literature it was claimed that the color-magnetic part of the OGE
itself is responsible for the repulsive core of the N-N interaction
\cite{barn}. We shall return to this point at the end of Section V.
The plot of $B$ as a function of $\beta$ is given in Fig. 1 for the
three cases (a), (b) and (c). $B$ converges towards a constant
value $\mu$, which can be calculated from considering two
well-separated non-interacting three-quark bags:
\begin{mathletters}
\begin{eqnarray}
\mu &=& 763.6 \mbox{MeV}\hspace{2cm} \mbox{in cases (a) and (b)} \ ,
\label{mue1}
\\
\mu &=& 502.3 \mbox{MeV}\hspace{2cm} \mbox{in case (c)} \ .
\label{mue2}
\end{eqnarray}
\end{mathletters}
We would expect $\mu$ to be equal to the reduced mass, $m_N$/2. The
discrepancy between the values of $\mu$ and $m_N$/2 -- which is
especially drastic if the OGE is included, i.e., in cases (a) and (b)
-- is related to the well-known Peierls-Yoccoz disease
\cite{PY57,RS80}.
\figa{fig1}{ht}{The effective mass, $B(\beta)$ of Eq.
(\protect\ref{mass}), as a function of the deformation parameter
$\beta$; the solid, dashed and dotted lines correspond,
respectively, to the cases (a), (b) and (c) introduced in Eq.
(\protect\ref{case}). The asymptotic values of Eqs.
(\protect\ref{mue1}) and (\protect\ref{mue2}) are indicated by
the arrows.}
\section{The Fujiwara transformation}
The dependence of the effective mass on $\beta$ prevents us from
directly interpreting the potential in Eq. (\ref{pot1}) as an ordinary
N-N potential. Moreover, $\beta$ doesn't correspond to the true N-N
separation distance (except for large positive $\beta$ when the two
nucleons are well separated). Therefore, we wish to transform Eq.
(\ref{schrod}) into a Schr\"{o}dinger-like equation with a constant,
coordinate independent mass term. For this purpose, one can use a
Fujiwara transformation \cite{Fu59,Wi89}, which relates the generator
coordinate $\beta$ to an effective N-N separation length:
\begin{equation}
x(\beta)=-\int^{\infty}_{\beta}\left[\sqrt{\frac{B(\beta')}{\mu}} -1
\right] d\beta' + \beta \ .
\label{fuji}
\end{equation}
If one now redefines the weight function in Eq. (\ref{schrod}) as
\begin{equation}
\tilde{\Phi}(\beta) = \root {\scriptstyle 4}
\of {\frac{B(\beta)}{\mu}}~ \psi(x) \ ,
\end{equation}
Eq. (\ref{schrod}) transforms into the familiar form
\begin{equation}
\left[-\frac{1}{2\mu}\frac{d^2}{dx^2}+V+V_F\right]\psi(x)=E\psi(x) \ ,
\end{equation}
with V given by Eq. (\ref{pot1}) and
\begin{equation}
V_F(\beta) = \frac{7}{32B^3}\left(\frac{dB}{d\beta}\right)^{\!2}
-\frac{1}{8B^2}\frac{d^2B}{d\beta^2} \ .
\label{potf}
\end{equation}
\figa{fig2}{bht}{The Fujiwara coordinate, $x(\beta)$ of Eq.
(\protect\ref{fuji}), as a function of the deformation parameter
$\beta$. The long-dashed line shows the asymptotic solution,
$x(\beta)=\beta$, and the remaining labeling is the same as in
Fig. 1.}
Figure 2 displays the explicit relationship between $x$ and $\beta$,
as obtained from Eq. (\ref{fuji}). As expected, the deformation
parameter $\beta$ converges asymptotically towards the effective
internucleon separation $x$. The correspondence
$\beta \leftrightarrow x$ should be
very useful in discussions of the quark substructure of nuclei or
nuclear matter using Schr\"odinger-based many-nucleon calculations
and employing our six-quark wave functions.
\section{Results for the effective N-N potential}
We now wish to present detailed results for:
\begin{equation}
V_{loc}(\beta) = V(\beta) + V_F(\beta) - V(\infty) \ ,
\label{sum}
\end{equation}
where $V(\beta)$ and $V_F(\beta)$ are given in Eqs. (\ref{pot1}) and
(\ref{potf}). The value of $V(\infty)$ corresponds to the asymptotic
value of $\tilde H_0/\tilde N_0$ calculated from two well-separated
non-interacting three-quark bags, and it is given in Section III.
This asymptotic value is the same in cases (a) and (b) because the
color-electrostatic mutual and self-energy terms cancel exactly due to
color neutrality when the two nucleons are well separated.
\figb{fig3a}{fig3b}{p}{The two contributions to the local potential,
$V_0(\beta)$ of Eq. (\protect\ref{v1}) and $V_1(\beta)$ of Eq.
(\protect\ref{v2}), as functions of the mean generator coordinate
$\beta$. The solid, dashed and dotted lines correspond
to the cases (a), (b) and (c) introduced in Eq.
(\protect\ref{case}).}
\figb{fig4a}{fig4b}{p}{The non-adiabatic, local potential, $V_{loc}$
of Eq. (\protect\ref{vloc}), as a function of the deformation
parameter $\beta$ and the Fujiwara coordinate $x$, respectively. The
labeling is the same as in Fig. 3.}
It is convenient to rewrite $V_{loc}(\beta)$ in the following form:
\begin{equation}
V_{loc}(\beta)=V_0(\beta)+V_1(\beta) \ ,
\label{vloc}
\end{equation}
with:
\begin{eqnarray}
V_0(\beta) &=& \frac{\tilde{H}_0}{\tilde{N}_0}-V(\infty) \ ,
\label{v1}\\
V_1(\beta) &=& \frac{1}{4B}\left[
\frac{d^2\ln\tilde{N}_0}{d\beta^2}
+\frac{1}{2}\left(\frac{d\ln\tilde{N}_0}{d\beta}\right)^{\!2}
-\frac{d\ln\tilde{N}_0}{d\beta}\frac{d\ln B}{d\beta}\right]
\nonumber\\
& &~~~+~\frac{1}{8B}\left[
\frac{3}{4}\left(\frac{d\ln B}{d\beta}\right)^{\!2}
-\frac{d^2\ln B}{d\beta^2}\right] \ .
\label{v2}
\end{eqnarray}
In order to calculate these derivatives, $\ln{B}$ and
$\ln{\tilde{N}_0}$ were fitted to polynomials.
The two contributions $V_i(\beta)$ (i=0, 1) to $V_{loc}(\beta)$
are plotted in Fig. 3 as functions of the deformation
parameter $\beta$ for the three cases outlined previously. Fig. 4
shows $V_{loc}$ as a function of $\beta$ and of the
Fujiwara coordinate $x$, respectively. Note that Eq. (\ref{v2}) was
obtained from Eq. (\ref{pot1}) by replacing $H_2 - E N_2$ with
$-\tilde N_0/B$, as indicated in Eq. (\ref{mass}).
The shape of $V_0(\beta)$ is quite similar to the adiabatic potentials
displayed in Fig. 10 of Ref. \cite{KW94}, both for the ``full OGE" and
``no-OGE" cases. This tends to confirm our
assumption that the matrix elements $\langle\,\beta+\delta/2\,|
\,H-E\,|\,\beta-\delta/2\,\rangle$ are rather sharply peaked around
$\delta=0$. The term $V_1(\beta)$ corresponds to the contribution of
non-adiabaticity. It grows important only for $\beta \lesssim -2$ fm,
and yields in all cases a repulsion due to the dynamics. This is
according to our expectation and in agreement with Ref.\cite{Sc86}.
Note that
in cases (b) and especially (c), we also obtain an intermediate range
attraction in $V_{loc}$.
The fact that our N-N potential extends to negative $x$ should not be
taken too literally. It simply reflects inadequacies in the
relationship between the deformation parameter $\beta$ and the N-N
separation length $x$, which are connected to the Peierls-Yoccoz
disease mentioned earlier.
We recall that one of the main objectives of this and our previous
study \cite{KW94} was to incorporate explicitly one-gluon exchange
effects, in contrast to Ref.\cite{Sc86} where they were neglected.
Comparing, for instance, cases (a) and (c), one can see that the OGE
reinforces the repulsive core considerably. The existence of a
repulsive core in all three cases makes us to attribute it to
dynamics rather than to the color-magnetic interaction (case (b)), as
was inferred in Ref.\cite{barn}.
\section{Conclusions}
In this investigation, we found that the dynamics are manifestly
responsible for the hard-core repulsion of the short-range part of the
N-N interaction, and we observed that we could obtain both short-range
repulsion and some intermediate range attraction if the entire
one-gluon exchange or at least its electrostatic part were neglected.
In the results containing the full OGE effects the lack of
attraction
is due to the omission of explicit meson exchanges. Then, to reproduce
the experimental phase shifts or other two-body data one necessitates
to attach a local OBE potential beyond a certain internuclear distance
\cite{Oka83}. To obtain this potential in the framework of our model
we could consider extending our calculations by either including
quantum surface fluctuations and/or introducing configurations of the
form $q^7\bar{q}$ in addition to the $q^6$ states. This would be a
rather cumbersome procedure within the present model. The most
convenient would be to either allow mesonic degrees of freedom and
to consider, e.g., an explicit pion exchange between the individual
quarks \cite{myhr} or to simply choose a phenomenological potential.
Another important result of this work is the evaluation of the
relationship between the deformation parameter $\beta$ and the
effective N-N separation length $x$ through the Fujiwara
transformation. This correspondence is very useful for applications
of our model to the description of phenomena involving the quark
substructure of light nuclei. It furthermore allows us to relate
many-body correlation functions or N-N wave functions given in the
literature to the GCM formalism presented here.
An attractive way to confirm our results would be to solve directly
the Hill-Wheeler integral equation in order to obtain phase shifts.
Projection on good angular momentum states should also improve our
calculations.
\acknowledgments{This work was supported in part by the MINERVA
Foundation of the Federal Republic of Germany, and by the U.S.
Department of Energy.}
|
\section{Introduction}
\label{s:intro}
Models of {\it random sequential adsorption} (RSA) have been used to describe
the process by which particles are irreversibly deposited without overlap
onto a surface.
They are relevant to studying the adsorption of gas molecules \cite{book} or
colloid particles in solution \cite{gas.diffusion} onto solid surfaces,
or of large molecules on biological membranes \cite{membranes}.
One of the key assumptions, is that the particles bind strongly to the
substrate so that desorption and surface mobility are negligible on the
time-scale of the experiment.
Although providing a very simplified picture, RSA models have the virtue
of being exactly solvable in dimension $d=1$, for monomer and for
$k$-mer particles \cite{1dsolution}.
Many additional features have been added to this model in order to make it
more physical, or to generalize its range of applications, and the literature
in the field is vast \cite{review}.
The inclusion of diffusional relaxation on the substrate
\cite{surf.diff,surf.diff-desorp}, leading to equilibration, or of cooperative
effects, such as multi-site exclusion \cite{nn.exclusion}, have been examined.
The possibilities that the particles are reflected back to the fluid or
desorbed \cite{surf.diff-desorp,HTW,King}, or that the surface comprises
more than one chemical species \cite{XK} have also been considered.
Another important possibility, motivated by many experimental studies of
gas-metal surfaces \cite{experiment,KW}, is that if the particles lose just
enough of their kinetic energy they can become trapped (physisorbed) in a
mobile, temporary {\it precursor state}, from which they can be adsorbed
(chemisorbed) at a far site at a later time.
The mechanism of precursor mediated chemisorption, first postulated by Taylor
and Langmuir \cite{precursor}, was initially formulated as a statistical model
by Kisliuk \cite{Kisliuk}. It was later adapted by King and coauthors
\cite{King,KW,CK,book} to include other effects, such as temperature
dependence,
desorption, molecular dissociation and pair interactions between the adatoms.
Variations on the Kisliuk model have also been studied by other authors, both
numerically and analytically \cite{HTW,XK,WBH,AD,BB,Evans,EN}.
Most analytical treatments, however, are essentially mean-field like, in that
they largely ignore the correlations between the state of occupation of
different sites which arise through the precursor mediated deposition process,
i.e.\ they assume that the rate of deposition at a given site only depends on
the state of occupation of that site, and not on the neighbouring sites.
In this paper, we consider a lattice model which is both a simplified version
of the Kisliuk model, and a slight complication of the RSA model.
Contrary to most previous studies, we examine the case where the mobility of
physisorbed atoms is only possible on the top of occupied chemisorption sites
({\it extrinsic} precursor states), and where there is neither reflection nor
desorption back to the gas phase (later we shall comment on how scattering and
desorption can in principle be included into the equations of motion).
Each deposition attempt will, therefore, result either in the direct occupation
of an empty site, or in the diffusion on the top of an island of occupied sites
until the particle finds an empty site at the edge of the island, where it is
irreversibly deposited.
For simplicity, we also consider that the deposited particles are
{\it monomers}, or atoms, in which case the lattice will eventually become
full.
As a consequence of the (extrinsic) precursor diffusion, the edges of islands
are preferential sites for chemisorption and the growth of the larger islands
is favoured.
This introduces non-trivial correlations between sites, especially at
late-times
when diffusion is the dominant deposition mechanism, and makes the model
unsolvable even for the simplest case of monomer deposition on a {\it line}.
A model similar to the one adopted here has been examined by Becker
and Ben-Shaul in $2d$ \cite{BB}.
Their analysis, however, only applies to the early kinetics, as they treat the
islands as uncorrelated objects, completely neglecting the contribution of
island coalescence to the growth process.
Regarding the motivation for our study, we recall a somewhat unrealistic
feature of the RSA model, namely that once a deposition attempt fails, the
particle's position is 'randomized' before another adsorption attempt is made.
As the transport of particles to the surface is diffusive, one might expect
that a failed deposition attempt would be followed by another nearby adsorption
attempt \cite{no.randomize}.
Since the intrinsic precursor seems to play no role in this effect, the present
model may account for it, at least partially.
Furthermore, it has been previously suggested by Cassuto and King \cite{CK},
that a kinetic deposition model without intrinsic precursor states and with
negligible desorption, could well also explain the experimental data from some
gas-solid systems, such as hydrogen on tungsten.
Naively, one may also think that such a model could be able describe the slow
deposition of liquid droplets on a plane surface, the spread of epidemics from
immunized to non-immunized populations, or the growth of a forest where seeds
have to be transported to an open field to find suitable conditions to develop.
Although island formation and structure kinetics have been studied
(see e.g.\ \cite{nn.exclusion,EN}), most work on RSA and precursor models,
either analytical or simulation, has focused on determining the dependence of
the coverage fraction $\theta$ and sticking probability on the exposure,
i.e.\ on the time $t$, and its value at the {\it jamming limit} (which is
trivial only for monomers).
In the present model, we assume that the diffusion time scale is small enough
compared to the time scale of deposition attempts, and so the coverage
is proportional to time or, equivalently, the sticking probability is 1.
We examine the case of a substrate of dimension $d=1$, and concentrate on
studying the quantities describing the evolving morphology of the occupied
regions. Namely, the pair correlation functions, the total number of
islands, and the probability distribution of the island sizes and its moments.
We have measured these quantities in numerical simulations, and developed a
minimal theory incorporating correlations, to calculate them approximately.
The results show good agreement, while at the same time differing considerably
from those of the RSA model for monomer deposition in $1d$ where there are no
spatial correlations at all.
A brief summary of this paper is the following.
In section 2 we describe the model and define the formalism.
The results of the numerical simulation are discussed in section 3.
The theory is presented in section 4, and the results are compared
with the simulation data. Section 5 is dedicated to some conclusions.
\section{Model}
\label{s:model}
We consider a one dimensional lattice with $N$ sites, with $N$ large enough
to neglect boundary effects, i.e.\ we take $N\to\infty$ in the calculations.
At each site $i$ we define a variable $S_i$, such that
\begin{eqnarray}
S_i & = & 0 \ \ \ {\rm if \ the \ site \ is \ empty} \nonumber \\
& = & 1 \ \ \ {\rm if \ the \ site \ is \ occupied} \nonumber \ .
\end{eqnarray}
Every time step $\delta t=1$ a site is chosen randomly and a monomer
deposition attempt is made.
If the site is empty the particle is adsorbed and the site becomes
occupied irreversibly.
Otherwise, the particle diffuses on top of the occupied region until it
reaches an empty site where it is adsorbed (this is appropriately modelled
by a random walk with traps \cite{random.walk}). Only then is the next
deposition attempted. The process repeats until the lattice is full.
We assume here that the particle diffusion is rapid enough to be over before
the time of the next deposition attempt, independently of the size of
the island where it takes place.
This may sound as an unrealistic assumption, especially if the island is of the
order of the system size. We note, however, that due to the randomness of the
process it is unlikely that a diffusing particle would interact or compete with
the next particle to be deposited. This argument fails, of course, in the limit
when the lattice is almost full (intermediate coagulation may occur) or in the
case when diffusion is very slow (a gas of net precursor particles develops).
Since each time step results in a deposition, either direct or mediated by
diffusion, the time dependence of the adsorbate coverage,
$\theta=\left<S_i\right>$, is trivial in this case:
\begin{equation}
\theta(t) = t/N \ \ , \ \ 0\leq t \leq N \ . \label{cov.D=1}
\end{equation}
One can easily incorporate in the model the possibility of {\it scattering}
of a particle back into the gas phase at the instant of collision with
the surface. If the scattering probability is $(1-\alpha)$, whatever
the state of occupation of the site, the rate of adsorption per unit time
must be multiplied by $\alpha$. We may then set $\alpha =1$ through a
redefinition of the time scale. To account for the possibility that scattering
may depend on the state of occupation of the site, we set the probability of
no chemisorption to be $(1-D\alpha)$ if a particle lands on an occupied site.
In this case, the rate of variation of the number of occupied sites is
(setting $\alpha=1$)
\begin{equation}
Nd\theta/dt = 1(1-\theta) + D\theta = 1 - (1-D)\theta \ , \label{cov.eq}
\end{equation}
with initial condition $\theta(0)=0$. The solution is
\begin{equation}
\theta(t) = \frac{1-\exp\left[-(1-D)t/N\right]}{1-D} \ . \label{cov.D}
\end{equation}
This reduces to (\ref{cov.D=1}) for $D=1$, and to the RSA behaviour for
$D=0$ and for $t\ll N$.
When $0<D<1$ the full coverage $\theta=1$ is attained at $t=N\log(1/D)/(1-D)$,
larger than $N$, due to the larger scattering from the occupied regions.
We may also account for {\it desorption}, simply by adding an extra negative
term (linear in $\theta$) to the rate equation (\ref{cov.eq}), effectively
decreasing the value of $D$.
The coefficient of the term may incorporate the probability of desorption
from the physisorbed or from the chemisorbed states, or both.
Allowing for desorption from the chemisorbed states would not only affect the
coverage, however, but will also imply extra terms in the rates of growth and
coalescence (eq.\ (\ref{R})).
For convenience we shall keep $D\neq 1$ in the future expressions, as it
allows us to distinguish the diffusion from the direct deposition terms.
Next we define several quantities which will be useful to examine the
morphology of the system. We shall call:
\begin{itemize}
\item $N(t)$ \ the total number of islands of occupied sites \\
\item $N(L)=N(0L0)$ \ the number of islands with $L$ sites, or number of rows
of $L$ consecutive occupied sites with at least one empty site on the left and
right \\
\item $N(L00)=N(0L00)$ \ the number of islands of $L$ sites with at
least two empty sites on the right \\
\item $N(L0L')=N(0L0L'0)$ \ the number of islands of $L$ sites separated
by an empty site from an island of $L'$ sites on the right \\
\item $N(00)$ \ the number of pairs of empty sites \\
\item $N(000)$ \ the number of trios of empty sites .
\end{itemize}
The generalization of the notation for more complicated configurations is
obvious.
The densities $n$ associated with the above numbers, are defined by
their ratio to the system size, e.g.\ $n(t)=N(t)/N$.
Since the boundary effects can be neglected, the number of occupied regions
(islands) is equal to the number of empty regions.
On average the system has left and right symmetry, so $N(L00)=N(00L)$
and $N(L0L')=N(L'0L)$.
For a $1d$ substrate, the above quantities are easily defined in terms
of the correlation functions of the local variables.
Although the system is not in equilibrium we expect it to be translationally
invariant, on average, due to the randomness of the deposition attempts.
Using this property we have, for example:
\begin{eqnarray}
n(t) & = & \sum_i \left<(1-S_i)S_{i+1}\right>/N
\ = \ \theta - \left<S_iS_{i+1}\right> \label{nt.1} \\
n(L) & = & \sum_i
\left<(1-S_i)S_{i+1}...S_{i+L}(1-S_{i+L+1})\right>/N \nonumber \\
& = & \left<(1-S_0)S_1...S_L(1-S_{L+1})\right> \label{nL.1} \\
n(L00) & = & \sum_i
\left<(1-S_i)S_{i+1}...S_{i+L}(1-S_{i+L+1})(1-S_{i+L+2})\right>/N \nonumber \\
& = & \left<(1-S_0)S_1...S_L(1-S_{L+1})(1-S_{L+2})\right> \label{nL00.1} \\
n(00) & = & \sum_i \left<(1-S_i)(1-S_{i+1})\right>/N
\ = \ 1 - 2\theta + \left<S_0S_1\right> \label{n00.1} \\
n(000) + n(1) & = & \sum_i \left<(1-S_i)(1-S_{i+2})\right>/N
\ = \ 1 - 2\theta + \left<S_0S_2\right> \label{n000.1}
\end{eqnarray}
A number of normalization or hierarchical sum rules follow, some of which are:
\begin{eqnarray}
N(t) & = & \sum_{L=1}^{\infty} N(L) \label{nt.sum} \\
N(L) & = & \sum_{L'=0}^{\infty} N(L0L')
\ = \ \sum_{L'=1}^{\infty} N(L0L') + N(L00) \label{nL.sum} \\
N(00) & = & \sum_{L=0}^{\infty} N(L00)
\ = \ \sum_{L=1}^{\infty} N(L00) + N(000) \label{n00.sum} \\
N(0) \ = \ (1-\theta)N & = & \sum_{L=0}^{\infty} N(0L0)
\ = \ N(t) + N(00) \label{n0.sum} \\
\theta N & = & \sum_{L=1}^{\infty} LN(L) \label{theta.sum} \ .
\end{eqnarray}
It is easy to show that expressions (\ref{nt.1}) and (\ref{nt.sum}) (with
$n(L)$ replaced by (\ref{nL.1})) are equivalent, using fixed or periodic
boundary condition and the property $S_i^2=S_i$.
The probability of finding an island of size $L$ and its moments,
the average island size and its mean square deviation, are therefore:
\begin{eqnarray}
P(L) & = & N(L)/N(t) \label{PL.1} \\
\left<L\right> & = & \sum_{L=1}^{\infty} LP(L) \ = \ \theta/n(t) \label{L.1} \\
\frac{\left<L^2\right>-\left<L\right>^2}{\left<L^2\right>} & = &
1 - \theta^2/\left(n(t)\sum_{L=1}^{\infty}L^2n(L)\right) \label{LL.1} \ .
\end{eqnarray}
There are three basic mechanisms by which these numbers may change with the
deposition process (where $\Delta$ denotes variation):
\begin{itemize}
\item {\it nucleation}: $\Delta N(000) = 1$ \\
\item {\it growth}: $\Delta N(L00) = - 1 \ ; \ \Delta N(L+1) = 1$ \\
\item
{\it coagulation}: $\Delta N(L0L') = - 1 \ ; \ \Delta N(L+1+L') = 1$ \ .
\end{itemize}
The rates of occurrence of each of these events per unit time are:
\begin{eqnarray}
R_n & = & \frac{N(000)}{N} \ = \ n(000) \nonumber \\
R_g & = & \sum_{L=1}^{\infty}
\frac{1+DL/2}{N}\left[N(L00)+N(00L)\right] \nonumber \\
& = & 2(n(00)-n(000)) + D\sum_{L=1}^{\infty}Ln(L00) \label{R} \\
R_c & = & \sum_{L=1}^{\infty}\sum_{L'=1}^{\infty}
\frac{1+D(L+L')/2}{N}N(L0L') \nonumber \\
& = & n(t)-n(00)+n(000)+D\theta-D\sum_{L=1}^{\infty}Ln(L00) \nonumber \\
& = & 1-(1-D)\theta-2n(00)+n(000)-D\sum_{L=1}^{\infty}Ln(L00) \nonumber \ ,
\end{eqnarray}
where we have used (\ref{nt.sum})-(\ref{theta.sum}).
The factor $L/2$ accounts for a particle landing and diffusing on top of
an island of size $L$ towards the right (or left).
For simplicity, we have assumed, and we will assume throughout, that in
modeling the diffusion process the random walk with traps can be replaced
by a random choice between right and left.
We shall return to this point in section \ref{s:simul}.
As expected, the sum of the rates (\ref{R}) yields the total
rate of adsorption per unit time: $\sum_i R_i=1-(1-D)\theta=Nd\theta/dt$.
The probabilities of nucleation, growth and coagulation per deposition event,
$P_n$, $P_g$ and $P_c$, are then defined by the ratios:
\begin{equation}
P_i=R_i/(Nd\theta/dt) \label{Pi} \ .
\end{equation}
It is then straightforward to write the exact equation for the rate of
variation of the number of islands, given by the difference between
the rates of nucleation and coagulation:
\begin{eqnarray}
\frac{dN(t)}{dt} & = & R_n - R_c \nonumber \\
& = & 1 - (D+1)\theta - 2n(t) + D\sum_{L=1}^{\infty} L\,n(L00) \label{nt.eq} \
,
\end{eqnarray}
where we have used (\ref{n0.sum}).
Setting $D=0$ in (\ref{nt.eq}) one obtains the RSA result
$n(t)=\theta(1-\theta)$.
Using (\ref{L.1}) and (\ref{cov.eq}), it is possible to derive from
(\ref{nt.eq}) an exact equation for the average island size $\left<L\right>$.
It is also illustrative to look at the growth of a single, isolated
island. Neglecting coagulation, the equation for the island size reads:
$dL/dt=(2+DL)/N$, which solution (with $L(0)=0$) is $L(t)=(2/D)(\exp(Dt/N)-1)$.
This, however, only gives the expected growth at early times.
When island coagulation becomes important the growth of $L$ should be faster
than exponential, and as $t\to N$ (for $D=1$) $L$ should become of order $N$.
\section{Simulation}
\label{s:simul}
We performed numerical simulations of the model described in section
\ref{s:model}. For convenience we used free boundary conditions, although the
choice of boundary conditions should be irrelevant.
The results were averaged over an ensemble comprising a great number ($N_s$) of
system samples with different random deposition histories.
For large coverages, the distribution $P(L)$ and its second moment proved
particularly sensitive to finite sampling and size effects.
This is easily understood, as the size of the larger island at late-times
(which controls the dynamics at this stage) can fluctuate by as much as $N/2$.
The smoothness of the curves for the probabilities $P_n$, $P_g$ and $P_c$ also
depends on the number of samples used.
We used $N=N_s=50000$, and verified that the systematic finite sampling and
size
errors were almost eliminated as we increased $N$ and $N_s$ up to these values.
For ease of analytical treatment (eq.\ \ref{R}) and for computational
efficiency, we modelled the precursor layer diffusion with a random
choice between right and left rather than with a random walk with traps.
Although that mechanism does not take into account the starting point of the
diffusion process, we expect this effect to be irrelevant because on average
all sites on an island are equally likely to be chosen for a deposition
attempt.
\subsection{Results}
\label{s:simres}
The measurements of $n(t)$, $1/\left<L\right>$,
$(\left<L^2\right>-\left<L\right>^2)/\left<L^2\right>$ and $P(L)$ (at six
different coverages) are displayed in Figures 1 to 4 (the pair correlations
are shown in section \ref{s:thres}).
We have divided the second moment by $\left<L^2\right>$ rather then
$\left<L\right>^2$ since the latter diverges.
It is interesting to observe the extent to which the RSA kinetics of deposition
is modified by the additional diffusion mechanism.
For comparison, we included in Figures 1 to 3 the plots (broken lines) of:
$\theta(1-\theta)$, $(1-\theta)$ and $\theta/(1+\theta)$, and in Figure 5 the
plot of $\theta^{L-1}(1-\theta)$, which correspond to the same quantities
in the RSA model.
We see from Figures 1 to 3, that the number of islands is smaller, and the
average island size and island size fluctuations are larger with diffusion
mediated deposition. This results from the increase in the rates of
island growth and coagulation relative to the rate of island nucleation.
Comparing Figures 4 and 5 for the island size distribution, we can see a
close agreement at small coverages, when all islands are still small.
At intermediate coverages, the distribution spreads out (larger islands)
in the diffusion case. At large coverages the difference is even
greater. The RSA curve becomes uniform (all island sizes are equally
probable), while the diffusion curve shows that the majority of islands
are still small in size, but there is a minority of very large islands.
At the late stages of diffusion mediated growth, the number of large islands
is small and their size can fluctuate enormously, therefore the
smoothness of the $P(L)$ curve depends strongly on the sampling number.
We have also examined the mechanisms responsible for the structure kinetics.
Figure 6 shows the evolution of the probabilities of nucleation, growth
and coagulation per deposition event, $P_n$, $P_g$ and $P_c$ (eq.\ (\ref{Pi})).
For comparison we have included the corresponding RSA curves (broken lines):
$P_{n,RSA}=(1-\theta)^2$, $P_{g,RSA}=2\theta(1-\theta)$ and
$P_{c,RSA} = \theta^2$ (cf.\ (\ref{R})-(\ref{Pi}) with $D=0$).
The most obvious difference occurs as $\theta\to 1$.
In the RSA case the probability of growth vanishes. In the diffusion case,
however, the probability of growth tends to a finite value $1-P_c$, showing
that a considerable number of gaps with two or more empty sites still exist
at large coverages.
\section{Theory}
\label{s:theor}
Due to the diffusion on top of the occupied regions, the probability of
adsorption at an empty site depends on the state of occupation of its
neighbouring sites and even of far located sites, if the site is at the edge
of a large island.
These correlations develop in the system as the coverage increases with
time and diffusion mediated deposition becomes more likely.
As a result, the density numbers defined in section \ref{s:model}
obey an infinite set of hierarchical coupled equations which cannot be solved
exactly (an example of an equation in the top of the hierarchy is given by
(\ref{nt.eq})). One must, therefore, look for approximate solutions by
truncating the hierarchy with some closure scheme.
The simplest approximation consists in incorporating only the {\it pair
correlations} in the system, which are then determined self-consistently.
We shall see that, by carefully choosing the multi-site functions decoupling
scheme, such a simple approach, despite neglecting multiple correlations,
is capable of capturing many qualitative and quantitative features of the
system's behaviour.
Let us denote the pair correlations between sites with different separations,
as:
\begin{eqnarray}
p \ = \ q^{(1)} & = & \left<S_iS_{i+1}\right> \nonumber \\
q \ = \ q^{(2)} & = & \left<S_iS_{i+2}\right> \label{pq.def} \\
q^{(n)} & = & \left<S_iS_{i+n}\right> \ , \ n\ge 1 \nonumber \ .
\end{eqnarray}
We expect the correlations to decay with distance, i.e.:
\begin{equation}
p \ge q \ge q^{(3)} \ge ... \ge \theta^2 \ . \label{decay}
\end{equation}
The quantities depending on the local variables of {\it two sites} only,
can be expressed exactly in terms of the pair correlations.
{}From (\ref{nt.1})-(\ref{n000.1}) we have:
\begin{eqnarray}
n(t) & = & \theta - p \label{nt.2} \\
n(00) & = & 1 - 2\theta + p \label{n00.2} \\
n(000) + n(1) & = & 1 - 2\theta + q \label{n000.2} \ ,
\end{eqnarray}
and the average island size (\ref{L.1}) is given by
\begin{equation}
\left<L\right> = \frac{\theta}{\theta-p} \label{L.2} \ ,
\end{equation}
To write the {\it multi-site} functions approximately, in terms of the pair
correlations, we decouple the higher order correlators into a product chain
of pair correlators, each associated with an adjacent bond, as follows:
\begin{eqnarray}
\left<S_1S_2...S_nS_{n+m}\right> & \simeq & \frac{\left<S_1S_2\right>
\left<S_2S_3\right>...\left<S_{n-1}S_n\right>\left<S_nS_{n+m}\right>}
{\left<S_2\right>...\left<S_{n-1}\right>\left<S_n\right>} \nonumber \\
& = & \frac{p^{n-1}q^{(m)}}{\theta^{n-1}} \ \ , \ \ (n\geq 2, m\geq 1) \ .
\label{pair}
\end{eqnarray}
The normalization factors in the denominator assure that the RSA result is
recovered in the decoupling limit. Then, from
(\ref{nL.1})-(\ref{n000.1}) we have:
\begin{eqnarray}
n(L) & \simeq & \frac{p^{L-1}}{\theta^{L}}(\theta-p)^2 \ , \ (L\geq 1)
\label{nL.2} \\
n(000) & \simeq & 1 - 3\theta + 2p + q - p^2/\theta \ . \label{n000.3} \\
n(L00) & \simeq & \frac{p^{L-1}}{\theta^{L+1}}(\theta-p)
\left[\theta^2-\theta(p+q)+p^2\right] \ , \ (L\geq 1) \label{nL00.2}
\end{eqnarray}
Note that the limits of ($n(00)-n(L)$) and ($n(000)-n(L00)$) when $L\to 0$,
which are $(p-\theta^2)/p$ and $(p-\theta^2+\theta(q-p))/p$, respectively,
although non-zero have a small value.
Using these expressions and (\ref{nt.2})-(\ref{n00.2}), we can then
write the deposition rates (\ref{R}) as:
\begin{eqnarray}
R_n & = & 1 - 3\theta + 2p + q - p^2/\theta \nonumber \\
R_g & = & (D+2)(\theta-p) + 2(p^2/\theta-q) + D\theta\frac{p-q}{\theta-p}
\label{R.appx} \\
R_c & = & Dp + q - p^2/\theta - D\theta\frac{p-q}{\theta-p} \nonumber \ ,
\end{eqnarray}
and the mean square deviation of the island sizes (\ref{LL.1}) as:
\begin{equation}
\frac{\left<L^2\right>-\left<L\right>^2}{\left<L^2\right>} \ = \
\frac{p}{\theta+p} \ . \label{LL.2}
\end{equation}
For the theory to be self-consistent, we must check that the expressions
for the densities, although approximated are properly normalized.
In fact, summing (\ref{nL.2}), with $L$ from 1 to $\infty$, gives $\theta-p$
(cf.\ (\ref{nt.sum}) and (\ref{nt.2})). Also, the sum
$\sum_{L=1}^{\infty}Ln(L)$
gives $\theta$ (cf.\ (\ref{theta.sum})).
Hence, the probability distribution for the island sizes (\ref{PL.1}),
obtained by dividing (\ref{nL.2}) by $n(t)$,
\begin{equation}
P(L) = (p/\theta)^{L-1}\left(1-p/\theta\right) \ , \label{PL.2}
\end{equation}
is normalized to 1.
This is a geometric distribution, as in the RSA case, but with $\theta$
replaced by $p/\theta$.
Adding (\ref{n000.3}) and (\ref{nL00.2}), with $L$ from 1 to $\infty$, gives
$1-2\theta+p$ (cf.\ (\ref{n00.2}) and (\ref{n00.sum})).
Other more complicated density numbers turn out to be consistently normalized
too, as their sums yield the correct density number within the pair
approximation. This is the case for $n(L0L')$, which satisfies the sum
rule (\ref{nL.sum}).
The next step of our approach is to determine the pair correlations
self-consistently.
There is an infinite hierarchy of coupled equations for the pair correlations,
even within the pair approximation. The first two are the equations for $p$ and
$q$. We will now derive the exact form of these two equations.
The $p$ equation follows immediately from equation (\ref{nt.eq}) for $n(t)$,
using (\ref{nt.2}) and (\ref{cov.eq}). It is more instructive, however,
to write it down by inspection.
$pN$ is the average number of pairs $11$, of two neighbouring occupied sites,
which increases by one with island growth and by two with coalescence. Hence,
using (\ref{R}), we have:
\begin{eqnarray}
N\frac{dp}{dt} & = & 2R_c + R_g \nonumber \\
& = & 2(D+1)\theta - 2p - D\sum_{L=1}^{\infty}Ln(L00) \ . \label{p.eq.exact}
\end{eqnarray}
Within the pair approximation, using (\ref{R.appx}) and (\ref{cov.D=1}), and
setting $D=1$, we obtain:
\begin{equation}
\frac{dp}{d\theta} = 3\theta - p - \theta\frac{p-q}{\theta-p} \ .
\label{p.eq.appx}
\end{equation}
The exact equation for $q$, although more complicated can also be written
down by inspection.
$qN$ is the average number of pairs $1-1$, of two occupied sites separated by
a site in any state.
It can increase by 1 or 2 with island nucleation, growth or coalescence.
A careful consideration of all possibilities leads to the following equation:
\begin{eqnarray}
N\frac{dq}{dt}
& \!\!\!\! & = \ 2\sum_{LL'=1}^{\infty}\frac{1}{N}N(L0{\bf 0}0L') +
2\sum_{LL'=2}^{\infty}\frac{1+D(L+L')/2}{N}N(L{\bf 0}L') \label{q.eq.exact1} \\
+ & \!\!\!\! &
2\sum_{L=1,L'=2}^{\infty}\frac{1+DL'/2}{N}
\left[N(L0{\bf 0}L')+N(L'{\bf 0}0L)\right] \nonumber \\
+ & \!\!\!\! & \sum_{L=1}^{\infty} \left\{
\frac{1}{N}\left[N(L0{\bf 0}00)+N(00{\bf 0}0L)\right] +
\frac{1+D/2}{N}\left[N(L0{\bf 0}1)+N(1{\bf 0}0L)\right] \right\} \nonumber \\
+ & \!\!\!\! & \sum_{L=2}^{\infty} \left\{
\frac{1+DL/2}{N}\left[N(L{\bf 0}00+N(00{\bf 0}L)\right] +
\frac{1+D(1+L)/2}{N}\left[N(L{\bf0}1)+N(1{\bf0}L)\right] \right\} \nonumber \ .
\end{eqnarray}
Employing the hierarchical relations (\ref{nt.sum})-(\ref{n0.sum}) and
some obvious generalizations, this equation then simplifies to
\begin{equation}
N\frac{dq}{dt} = 2 - 2(1-D)\theta - (2+D)n(1) - 2n(000) -
D\sum_{L=1}^{\infty}L\left[n(L000)+n(L010)\right] \ . \label{q.eq.exact2}
\end{equation}
Evaluating the sum of $L[n(L000)+n(L010)]$ within the pair approximation,
using (\ref{nL.2}) and (\ref{n000.3}) and putting $D=1$, yields
\begin{equation}
\frac{dq}{d\theta} = 2(\theta+p-q) -p^2/\theta +
\frac{\theta q^{(3)}-pq}{\theta-p} \ . \label{q.eq.appx}
\end{equation}
One can check that $n(L000)$ and $n(L010)$ are properly normalized within
the pair approximation (their sums give $n(000)$ and $n(1)$, respectively).
The approach also yields $\sum_{L=0}^{\infty}[n(L000)+n(L010)]=1-2\theta+q$,
which is the exact result (cf.\ (\ref{n000.2})).
It is straightforward to derive the behaviour of $p$ and $q$ for small and for
large coverage.
As $\theta\to 0$ the system is RSA like, so $p$, $q$ and $q^{(3)}$ should
behave as $\theta^2$. Equations (\ref{p.eq.appx}) and (\ref{q.eq.appx}) reduce
to $p' = 3\theta + O(\theta^2)$ and $q' = 2\theta + O(\theta^2)$, where
primes indicate derivatives with respect to $\theta$. Hence
\begin{eqnarray}
p & = & 3/2\, \theta^2 + O(\theta^3) \label{p.to0} \\
q & = & \theta^2 + O(\theta^3) \ \ , \ \ (\theta\to 0) \label{q.to0} \ .
\end{eqnarray}
The factor $3/2$ (confirmed by the simulations, Figure 7) shows that even in
this regime there is an increase in the correlations relative to the RSA case.
It results from the diffusion mediated growth of single site islands:
with RSA there 2 possibilities for growth, and with diffusion there is a
third one; therefore, there are $3/2$ as many double site islands as in RSA.
Let us now consider the limit when $\theta\to 1$.
At $\theta=1$, (\ref{p.eq.appx}) and (\ref{q.eq.appx}) yield the equations:
$(2-p'(1))(1-p'(1)) = p'(1)-q'(1)$ and
$(2-q'(1))(1-p'(1)) = q'(1)-q^{(3)'}(1)$.
The first equation and inequality (\ref{decay}) imply that
$1\leq p'(1)\leq q'(1)\leq q^{(3)'}(1)\leq ...\leq 2$. Hence, the solution is
\begin{eqnarray}
p'(1) & = & 2-(2-q'(1))^{1/2} \ = \ 2-(2-q^{(3)'}(1))^{1/3} \nonumber \\
q'(1) & = & 2-(2-q^{(3)'}(1))^{2/3} \label{pq.der} \ .
\end{eqnarray}
The actual value of the derivatives depends on the higher derivative
$q^{(3)'}$, and thus on the truncation scheme adopted.
The simulation results (Figure 7) suggest that $p'(1)=q'(1)=2$,
which implies (via (\ref{pq.der}) that $q^{(3)'}(1)=2$, and that
\begin{equation}
p \ = \ q \ = \ q^{(3)} \ = \ 2\theta - 1 + O(1-\theta)^2 \ = \
\theta^2 + O(1-\theta)^2 \ \ \ , \ \ \ (\theta\to 1) \label{pq.to1} \ .
\end{equation}
This is consistent with the truncation schemes: $q=\theta^2$ or
$q^{(3)}=\theta^2$.
\subsection{Results}
\label{s:thres}
The $p$ equation (\ref{p.eq.appx}) involves $q$;
the $q$ equation (\ref{q.eq.appx}) involves $q^{(3)}$, etc.
The nature of the approximation depends on how we close the hierarchy.
In the simplest, {\it first approximation} we neglect the correlations
beyond the nearest neighbours, i.e.\ we set $q=q_1=\theta^2$, solve numerically
equation (\ref{p.eq.appx}) for $p=p_1$, with initial condition $p(0)=0$, and
substitute $p_1$ and $q_1$ in the quantities of interest.
In the {\it second approximation} we neglect the correlations
beyond the second
neighbours, i.e.\ we set $q^{(3)}_2=q^{(3)}=\theta^2$, and
solve the system of equations (\ref{p.eq.appx})-(\ref{q.eq.appx}) for
$p=p_2$ and $q=q_2$, with initial conditions $p(0)=q(0)=0$.
Figure 7 shows the simulation data, $p_s$ and $q_s$, and the predictions
from the first theory, $p_1$ and $q_1=\theta^2$.
The correlations from the simulation decay with distance as in (\ref{decay}),
and the differences $p_s-\theta^2$, $q_s-\theta^2$ are small but non-zero,
as they should be since they establish the difference of behaviour
relative to the RSA model. The agreement between $p_1$ and $p_s$ is quite good.
As expected, it gets slightly worse as $\theta$ approaches 1 and the
correlations between sites further apart become more relevant, but the
correct asymptotic behaviour is obtained.
Consequently, there is also good agreement in the results for the density of
islands $n(t)$ (Figure 8) and the average island size $\left<L\right>$
(Figure 9), which are (exact) functions of $p$ only.
The agreement between $q_1$ and $q_s$ (Figure 7) is, of course, less
satisfactory.
Figure 10 compares the theoretical and simulation plots for the probabilities
of nucleation, growth and coagulation, $P_n$, $P_g$ and $P_c$
(eq.\ (\ref{R.appx}) with $D=1$; cf.\ Figure 6).
There is good quantitative agreement up to $\theta=0.5$, and
there is still some qualitative agreement for larger coverages.
The theory fails, however, to give the correct asymptotic behaviour as
$\theta\to 1$: although there is a region where $P_g\simeq 1-P_c$, the
theory gives $P_g\to 0$ as $\theta\to 1$.
{}From (\ref{R.appx}) we can see that $P_g\simeq (1-q')/(1-p')-1$ close to
$\theta=1$. Hence the limit $P_g=0$ is a consequence of having $p'(1)=q'(1)$,
which follows (via (\ref{pq.der}) from the truncation $q=\theta^2$.
Nearest-neighbour correlations are sufficient to probe the presence of island
boundaries. Since the above quantities depend essentially on island counting,
their predictions are fairly accurate.
Except in the early stages of deposition, however, long-range correlations need
to be accounted for to correctly describe the spectrum of island sizes.
Hence, as a result of the pair decoupling approximation (\ref{pair}), the
theory predicts a geometric distribution for the island size probability
(eq.\ (\ref{PL.2})), the same as in the RSA case but with $\theta$ replaced by
$p/\theta$.
The plots of $P(L)$ are analogous to the RSA ones (Figure 5), though since
$p/\theta>\theta$ each curve appears to correspond to a slightly larger
coverage.
The second moment of $P(L)$ (eq.\ (\ref{LL.2})) is, of course, also incorrect:
its limit as $\theta\to 1$ is $1/2$ as in RSA, rather than 1 as in
the simulation.
{}From the plots of $P(L)$ and its second moment, we find that the theory
breaks
down for these quantities for coverages over $0.2$.
Figure 11 compares the simulation data with the predictions from the
second theory, $p_2$, $q_2$ and $q^{(3)}_2$.
As before, $p_2$ fits $p_s$ quite well, and $q_2$ fits $q_s$ even better.
Consequently, the agreement in the results for $n(t)$ (Figure 8) and
$1/\left<L\right>$ (Figure 9) is also good: there is no major difference
between the two theories, apart from the fact that now the curves lay below
the simulation plots.
A similar difference between theories is found (Figure 10) for the probability
of nucleation $P_n=n(000)$, an approximate function of both $p$ and $q$
(eq.\ (\ref{n000.3})).
A worse agreement with the simulation than in the first theory is
obtained, however, for the probabilities of growth and coagulation.
We have also tested other plausible choices for the closure scheme, as
the ones employed above are not unique, but found the results were
either largely unchanged or incorrect.
Finally, we note that $p_2$ lies over $p_s$, while $p_1$ lies below $p_s$.
$q_2$, on the other hand, lies between $q_s$ and $q_1$.
Hence, the $p$ correlations are underestimated in the first theory
and overestimated in the second theory.
This seems to indicate that the approach cannot be systematically improved by
higher order truncations in the pair correlation hierarchy.
This fact comes as no surprise given the uncontrolled nature of a
self-consistent approach. Moreover, the results are more likely to be
affected by the pair decoupling approximation (\ref{pair}) then by the
order of truncation in the pair correlation hierarchy.
\section{Conclusions}
\label{s:concl}
We have studied numerically and analytically a simple, but non-trivial
model for the deposition of monomers on a line.
The particles can diffuse on the extrinsic precursor layer until they
reach the edge of the island, where they are irreversibly deposited.
As time progresses, island nucleation becomes less frequent, while the
larger islands grow rapidly and merge with other islands.
During this process, the precursor particles migrate for larger and larger
distances with ever increasing probability, establishing correlations
between sites further and further apart.
As a result, the system develops a structure characterized by strong
correlations whose range grows to the system size at full coverage.
In the simulation, we looked at the evolving structure pattern by
measuring the island density number, the island size probability distribution
and its first and second moments, and some of the pair correlations.
We also looked at the interplay between the direct and the diffusion
mediated mechanisms of deposition, by measuring the nucleation, growth
and coalescence probabilities per unit time.
As expected, the results (Figs.\ 1-7) differ considerably from the RSA model,
especially at large coverages, due to the increasing correlations.
To explain and interpret these measurements, we proposed a simple,
self-consistent theory which incorporates correlations to a minimum extent,
i.e.\ local pair correlations.
We considered two levels of approximation, depending on the closure of
the hierarchical equations for the pair correlations.
Altogether, the lowest level of approximation, accounting for
nearest-neighbour correlations only, gave the best fit to the simulation data.
The predictions of the theory (Figs.\ 7-9) proved very accurate for the
nearest-neighbour correlator $p$, island density number $n(t)$ and average
island size $\left<L\right>$.
A fairly good prediction was also obtained for the probabilities of nucleation,
growth and coagulation (Fig.\ 10).
While the nearest-neighbour correlations are sufficient to distinguish
occupied from non-occupied regions, the full range of correlations is
required to distinguish the sizes of those regions.
An accurate determination of the island size distribution $P(L)$ and
its second moment is, therefore, beyond the scope this theory, and
would, in principle, require the use of more complicated methods.
We end with a comment on the $2d$ systems, which are of more interest
to experimentalists.
In this case, however, the same quantities are not easily expressed
in terms of the local lattice variables, and one is faced with basic
difficulties in the development of a useful formalism (most of section
\ref{s:model} would not be applicable) and in the analytical treatment.
To illustrate the problem, we note that instead of the nearest-neighbour
correlations, some non-local operator would be required to probe the domain
boundaries.
It is desirable and possible, nonetheless, to perform numerical simulations
of the $2d$ model, which would also enable the study of richer phenomena,
such as percolating clusters.
\section*{Acknowledgments}
We would like to thank Adrian Taylor for valuable discussions at the
outset of this work.
\newpage
\baselineskip=12pt
|
\section{Introduction}
The problem of the Random Walk in presence of a boundary line near the
so--called {\it compensation point} for long chains has been solved years
ago by using standard statistical mechanics methods (see \cite{EKB} and
references therein). It is nevertheless worth to reconsider this model in
the light of recent developments in boundary Quantum Field Theory
\cite{goza}, in order to understand in a deeper way the connection between
the classical configurations of chains and Green's functions in the
corresponding Quantum Field Theory model. In this paper, we will show that
the statistical problem of the 2-d Random Walk with a boundary line can be
mapped onto a bosonic Quantum Field Theory with a defect line. Namely, we
will see that in order to derive the statistical behaviour of the Random
Walk in the presence of a boundary condition, one has to treat the
boundary not as a pure classical object but as a quantum defect line in
the corresponding free massive boson model, where both Reflection and
Transmission amplitudes are needed. As a by-product of our results, we
show that the sum of the aforementioned amplitudes plays
the role of the
boundary $S$--matrix for the
free massive bosonic Quantum Field Theory in
half--plane, such that a definition of a boundary state for this problem
can be used to compute the quantities we are interested in.
The Quantum Field Theory approach presented in this paper may be useful to
analyse the analogous problem with the Random Walk substituted by the Self
Avoiding Walk. We would like to remind that in the bulk, many
geometrical quantities of the Self Avoiding Walk can be obtained by using
an $S$--matrix approach \cite{Zam1,cm}, relying on the relationship
between Self Avoiding Walks and the $O(n)$ model for $n\rightarrow 0$
\cite{deg}. This relationship has already been used to discuss several
interesting aspects in the presence of a boundary condition\footnote{
Fendley and Saleur \cite{FS} have recently conjectured the exact boundary
$S$--matrix for the Self Avoiding Walk, by using an analogy with the
corresponding amplitude of the {\it Kondo problem}. It would be
interesting to have a direct derivation of this quantity as a solution of
the functional equations satisfied by the boundary $S$--matrix.}.
\section{ The Random Walk with boundary}
In this section, we will review some results of the Random Walk with
boundary in order to establish the correspondence with the language of
Quantum Field Theory. We closely follow the formulation given in
ref.~\cite{Eis} (for the problem in the bulk, see \cite{ItzD} and
\cite{MonWest}).
Let us initially consider the simplest model: the one--dimensional Random Walk
on the lattice, with the walker confined to move only on the positive
half--line $x\geq 0$. With a potential
\begin{equation}
V=\left\{ \begin{array}{ll}
\epsilon & \mbox{if $x=0$} \\
0 & \mbox{if $x\geq 1$},
\end{array}
\right.
\end{equation}
the partition function
for the configurations is given by
\begin{equation}
Z_V(x,x_0;N)=\sum_{n_0=0}^{\infty}\,a^{n_0}\, Z(x,x_0;N;n_0),
\label{parv}
\end{equation}
where $a\sim e^{-\epsilon/kT}$ and $n_0$ is the number of times a given path
{\it sits} in the origin. The partition function $Z$
in the rhs of (\ref{parv})
counts the number of different configurations, in the absence of potential,
of a chain of length $N$ with fixed ends ($x$, $x_0$) and which sits
$n_0$ times in the origin. By using the {\it images method} \cite{Cha}, this
expression can be reduced to
\begin{equation}
Z_V(x,x_0;N)= Z_b(x,x_0;N;n_0=0) +
2\sum_{n_0=1}^{\infty}\,\left(\frac{a}{2}\right)^{n_0}\,Z(x,x_0;N;n_0) \, ,
\label{para}
\end{equation}
where $Z_b$ is the partition function in the bulk.
For $\epsilon<0$, there exists a critical temperature $T_c$ such that for
$T=T_c$ we get $a_c=2$. This value of the temperature defines the
so--called {\it compensation point}, where the walker does not feel any
driving force, neither the (entropic) repulsion nor the (energetic)
attraction. In fact, for $T>T_c$, we observe a preference for the walker
to escape from the potential well, i.e. the favourite configurations are
those which end far away from the boundary. This will be called the {\it
non-adsorbed phase} of the Random Walk. On the contrary, for $T<T_c$ the
favourite configurations are those approaching the boundary with a low
probability to escape. This will be called the {\it adsorbed phase}.
The existence of two distinct phases of the Random Walk and a critical
point in between can be also established in the case of two--dimensional
Random Walk with boundary \cite{Eis}. In the continuum limit,
in order to mimic the boundary around
the {\it compensation point}, the potential can be chosen as
\begin{equation}
W=\left\{ \begin{array}{ll}
\infty & \mbox{if $x\leq 0$} \\
< 0 & \mbox{if $0< x < b$} \\
0 & \mbox{if $x\geq b$}
\end{array}
\right.
\end{equation}
and independent from the coordinate
parallel to the boundary line, say $y$. Since the two--dimensional
partition function of the Random
Walk can be factorized into the product of two independent one--dimensional
partition functions, we will study first the one--dimensional problem and
then we will come back to the original two--dimensional case.
The Green's function of the Random Walk, given by the Laplace transform of
the partition function $Z$, in the one--dimensional case is the solution
of the differential equation
\begin{equation}
\label{defG}
\left(-\partial^2_x + m^2 +W(x)\right)G(x,x_0;m^2)=\delta(x-x_0),
\end{equation}
with the additional condition that it vanishes at infinity.
The above differential equation can be solved by using standard methods
(see for example \cite{smir}). Here we concentrate our attention on the
solution given by
\begin{equation}
\label{grf}
G(x,x_0;m^2)=\frac{e^{-m|x-x_0|}+F(m,b,T)e^{-m(x+x_0)}}{2m}
\end{equation}
for $x, x_0\geq0$ where all informations about the potential are encoded
into the function $F$. This function can be cast into the following
universal form \cite{EKB} (see also \cite{deg1})
\begin{equation}
\label{F} F=\frac{1-c/m}{1+c/m}\, ,
\end{equation} provided that the
length of the chain $\sqrt{N}>>b$ and that the function $c\propto (T-T_c)$
satisfies\footnote{The lower bound for $c$ negative comes from the
requirement that the denominator of (\ref{F}) be bigger than zero. We are
supposing to have chosen a potential $W$ that satisfies these properties
\cite{Eis}.} $-b^{-1}<<c\leq b^{-1}$. Since we are interested in the
universal behaviour of the Random Walk chains, we may let at this point
$b\rightarrow 0$ and consider the Green's function (\ref{grf}) with the
above function $F$ as meaningful expressions for any $x,x_0\geq 0$. This
obviously implies that we are not concerned, from now on, with a microscopic
analysis of the interaction, much like in the spirit of the $S$--matrix
approach for the particle models.
We note here the following limits:
\begin{description}
\item[a)]
\noindent
for $c\rightarrow +\infty$, the function $F\rightarrow -1$
and the Green's function becomes
\[
G(x,x_0;m^2)=\frac{e^{-m|x-x_0|}-e^{-m(x+x_0)}}{2m} \,\,\, .
\]
This limit corresponds to the {\it hard--wall} behaviour for
$x,x_0$ far away from the boundary.
\item[b)]
\noindent
for $c\rightarrow 0$, we have instead
\[G(x,x_0;m^2)=\frac{e^{-m|x-x_0|}+e^{-m(x+x_0)}}{2m}\, .\]
This allows us the identification of the point $c=0$ in this description
as the {\it compensation point} of the Random Walk with boundary.
\end{description}
Finally, it is important to notice that the Green's function (\ref{defG}) of
the one--dimensional Random Walk can be also obtained as solution of the
{\it free} differential equation in one--dimension \cite{Eis}
\begin{equation}
\label{defGf}
\left(-\partial^2_x + m^2 \right)G(x,x_0;m^2)=\delta(x-x_0)\, ,
\end{equation}
but with the interaction encoded into the boundary condition
\begin{equation}
\label{boc}
\partial_x\,G(0,x_0;m^2)=c\,G(0,x_0;m^2)\, .
\end{equation}
Once the solution of the one--dimensional case has been obtained, the
Green's function of the two--dimensional Random Walk can be computed by
using Fourier transform as
\begin{equation}
G({\bf r},{\bf r}_0;m^2)=\int_{-\infty}^{+\infty}\frac{\mbox{d}k}{2\pi}\,
e^{ik(y-y_0)}\,G(x,x_0;m^2+k^2),
\end{equation}
where $G(x,x_0;m^2+k^2)$ satisfies the differential equation (\ref{defG}) and
is given by (\ref{grf}) and (\ref{F}) with the substitution
$m^2\rightarrow m^2+k^2$. This integral can be cast in the
suitable form
\begin{equation}
\label{gtot}
\begin{array}{c}
\begin{displaystyle}
G({\bf r},{\bf r}_0;m^2)=\int_0^{\infty}\frac{\mbox{d}\theta}{2\pi}\,
e^{im(y-y_0)\sinh\theta}\left(e^{-m|x-x_0|\cosh\theta} +\right.
\end{displaystyle}
\\
\\
\begin{displaystyle}
\left. +\hat{S}(\theta,m,c)\, e^{-m(x+x_0)\cosh\theta}\right)\, ,
\end{displaystyle}
\end{array}
\end{equation}
where
\begin{equation}
\label{bs}
\hat{S}(\theta,m.c) = \frac{\cosh\theta -c/m}{\cosh\theta +c/m}.
\end{equation}
{}From the equations (\ref{defGf}) and (\ref{boc}) satisfied by the
one--dimensional Green's function, it is simple to derive the differential
equation satisfied by the above one
\begin{equation}
\left(-\Delta_{\mbox{\bf r}} + m^2 \right)G(\mbox{\bf r},\mbox{\bf r}_0;m^2)
=\delta(\mbox{\bf r}-\mbox{\bf r}_0)
\end{equation}
supplied with the boundary condition
\begin{equation}
\left. \partial_x G({\bf r},{\bf r}_0;m^2)\right|_{x=0}=
\left. c\,G({\bf r},{\bf r}_0;m^2)\right|_{x=0}.
\end{equation}
Notice that the above equations are those satisfied by the two-point
correlation function for the euclidean massive boson with action
\begin{equation}
\label{action}
S[\varphi]=\int\,{\mbox d}x\,{\mbox{d}}y\,
\left\{\theta(x)\,
\left( \frac{1}{2}(\nabla\varphi)^2+\frac{m^2}{2}\varphi^2\right) +
\frac{c}{2}\delta(x)\varphi^2\right\},
\end{equation}
where $\theta(x)$ is the Heaviside distribution.
\section{The Quantum Field Theory approach}
Aim of this section is to show that there exists a one-to-one correspondence
between the problem of two--dimensional Random Walk with boundary near the
compensation point
and a Quantum Field Theory of a bosonic field $\varphi$
with a line of defect.
In particular, we will show that the pure {\it hard--wall}
situation in the Random Walk ($T\rightarrow\infty$) is described in terms
of a {\it totally reflective} defect in the Quantum Field Theory
model, while the {\it compensation point} ($T=T_c$) of the Random Walk
corresponds to a
{\it totally transmitting} defect, provided that the (classically) forbidden
negative half line is mirrored to the positive axis.
In order to establish this correspondence, the first step is to associate
to each chain of the Random Walk problem a trajectory of the particle
field $\varphi$ described by the Quantum Field Theory\footnote{In the context
of polymer
physics, this interpretation has been proposed in \cite{Zam1}.}. The second
step consists in solving a combinatorial problem arising from the counting of
the configurations. To this aim,
it will be convenient to consider two copies of the Random Walk problem,
defined on the left and right sides of the boundary respectively. The two
copies are subjected to the same potential well and share the same
temperature. In this picture, the boundary may be treated as a defect line.
Notice that, since at the {\it compensation point} the behaviour of the
Random Walk is like that in the absence of boundary, this corresponds, in the
two-copy scheme, to trajectories that start e.g. from the right side of the
boundary and end to the left side of it or viceversa. Said in other words,
the {\it compensation point} is mapped into the {\it pure transmitting}
behaviour of the defect line. Viceversa, the {\it hard--wall} limit
of the Random Walk corresponds to {\it purely reflecting} scattering
processes at the defect line.
Let us formulate more precisely this mapping. Consider the following action
\begin{equation}
S[\varphi_L,\varphi_R]=S[\varphi_L]+S[\varphi_R]
\label{action1}
\end{equation}
where
\[
S[\varphi_R]=\int\,{\mbox d}x\,{\mbox d}y\,\left\{\theta(x)\,\left(
\frac{1}{2}(\nabla\varphi_R)^2+\frac{m^2}{2}\varphi_R^2\right) +
\frac{c}{2}\delta(x)\varphi_R^2\right\}\]
and
\[S[\varphi_L]=\int\,{\mbox d}x\,{\mbox d}y\,\left\{\theta(-x)\,\left(
\frac{1}{2}(\nabla\varphi_L)^2+\frac{m^2}{2}\varphi_L^2\right) +
\frac{c}{2}\delta(x)\varphi_L^2\right\}.\]
The fields $\varphi_{L,R}$ are not independent since are linked each other
by the equation
\begin{equation}
\varphi_R(y,x)=\varphi_L(y,-x)\, .
\label{symc}
\end{equation}
The equations of motion associated to the action (\ref{action1}) are given by
\begin{equation}
\begin{array}{c}
\theta(x)\left(-\nabla_x + m^2\right)\varphi_R=0\\
\\
\theta(-x)\left(-\nabla_x + m^2\right)\varphi_L=0\\
\\
\left. \partial_x\left(\varphi_R-\varphi_L\right)\right|_{x=0}=
\left. c\,\left(\varphi_R+\varphi_L\right)\right|_{x=0}\\
\\
\varphi_R(y,0)=\varphi_L(y,0)
\end{array}
\end{equation}
where the last equality comes from equation (\ref{symc}).
Now we are in the position to see that this set of equations are the
euclidean version of those solved by A.J. Bray and M.A. Moore in \cite{BM},
who computed the Green's function (\ref{gtot}), and by G. Delfino,
G. Mussardo and P. Simonetti for the problem of the free relativistic
massive boson with a {\it line of defect} \cite{dms}. As proved in \cite{dms},
the dynamics of the massive boson with a line of defect is constrained by
the integrability conditions and is completely encoded into a set of
Transmission and Reflection amplitudes associated to the scattering processes
of the particle hitting the defect (Figure 1). Their explicit expressions are
given by \cite{dms}
\begin{equation}
\label{TR}
\begin{array}{c}
\displaystyle{
T(\beta,c)= \frac{\sinh\beta}{\sinh\beta+ic/m}}\\
\\
\displaystyle{
R(\beta,g)=-\frac{ic/m}{\sinh\beta+ic/m}}
\end{array}
\end{equation}
where now $\beta$ is the rapidity variable defined through the identity
\[ (E,p)=(m\cosh\beta,m\sinh\beta).\]
The remaining part of this letter will be devoted to the computation of the
two--point correlation function of the field $\varphi$ in the presence
of the defect line and to show that this quantity gives rise to the
Green's function (\ref{gtot}) of the Random Walk problem.
The correlation functions of the bosonic field $\varphi$ can be computed
by using the Form Factor approach for the integrable models \cite{smi,KW}.
This can be conveniently done by considering
the model defined in a geometry where the boundary or the defect are placed
at $t=0$. In this geometry, the boundary or defect line are promoted to
quantum operators which acts on the vacuum of the
bulk quantum theory whereas
the matrix elements of the fields remain those given in the bulk
case.
The Form Factors come from the insertion, in the correlation functions of a
given operator ${\cal O}({\bf r}),$ of a complete set of asymptotic states
\[
|\theta_1\cdots\theta_n>=A^{\dagger}(\theta_1)\cdots A^{\dagger}(\theta_n)|0>
\]
in such a way that for the time-ordered product of two such operators we have
\[
\begin{array}{l}
<0|{\cal O}({\bf r}_2){\cal O}({\bf r}_1)|0>=
\begin{displaystyle}
\sum_{n=0}^{\infty}\frac{1}{n!}\int\frac{{\mbox{d}}\theta_1
\cdots{\mbox{d}}\theta_n}{(2\pi)^n}|F_{{\cal{O}}}(\theta_1,\ldots,\theta_n)|^2
\cdot
\end{displaystyle}
\\
\\
\cdot \exp\left[im(y_2-y_1)\sum_{i=0}^n
\sinh\theta_i-im|t_2-t_1|\sum_{i=0}^n\cosh\theta_i\right].
\end{array}
\]
The Form Factors are defined by
\[
F_{{\cal{O}}}(\theta_1,\ldots,\theta_n)=
<0|{\cal O}(0)|\theta_1\cdots\theta_n>\, .
\]
In our case, the creation and annihilation operators of the particle states
satisfy the usual bosonic commutation relations
\[
\left[A(\theta),A^{\dagger}(\beta)\right]=2\pi\delta(\theta-\beta)\, ,
\]
and this drastically simplifies the calculation of the Form Factors. In
fact, for the fields $\varphi_{L,R}$ we have
\begin{equation}
<0|\varphi(0)|\theta_1\cdots\theta_n>=\frac{1}{\sqrt 2}\delta_{n,1}
\end{equation}
while all the other non--vanishing Form Factors can be
computed by using Wick's theorem based on the algebra of the operators
$A(\theta)$ and $A^{\dagger}(\theta)$. With the above matrix elements, the
euclidean correlation function in the bulk is given by
\begin{equation}
<0|{\cal T}\left[\varphi(y,t)\varphi(y_0,t_0)\right]|0>_E=
\int_0^{\infty}\,\frac{\mbox{d}\theta}{2\pi}\,e^{-m|x-x_0|\cosh\theta
+im(y-y_0)\sinh\theta}\, ,
\end{equation}
where in the rhs we have set $it=x$.
Let us now consider the problem of computing correlation functions in the
presence of the defect line.
By considering the
scattering processes as occur at the defect line in their crossed channels
(Figure 2.a, 2.b), we need the new amplitudes given by
\begin{equation}
\begin{array}{l}
\hat{T}(\theta,c)=T(i\frac{\pi}{2}-\theta,c)=
\displaystyle{
\frac{\cosh\theta}
{\cosh\theta+c/m}} \\
\\
\hat{R}(\theta,c)=R(i\frac{\pi}{2}-\theta,c)=\displaystyle{
\frac{-c/m}{\cosh\theta+c/m}}\, .
\end{array}
\end{equation}
The computation of the correlation functions in presence of the defect
operator ${\cal{D}}$ can be performed by using the equations
\[
<\varphi(y_1,t_1)\cdots\varphi(y_n,t_n)>=
\frac{<0|{\cal{T}}[\varphi(y_1,t_1)\cdots{\cal{D}}\cdots\varphi(y_n,t_n)]|0>}
{<0|{\cal{D}}|0>}
\]
where the matrix elements of the operator ${\cal D}$ are given by
\cite{dms}
\[
\begin{array}{c}
<\beta|{\cal{D}}|\theta>=2\pi\hat{T}(\beta,c)\delta(\beta-\theta)\\
\\
<\beta_1,\beta_2|{\cal{D}}|0>=2\pi\hat{R}(\beta,c)\delta(\beta_1+\beta_2)
\end{array}
\]
and $<0|{\cal{D}}|0>=1.$
There are two cases to consider: the first case is when the two operators
$\varphi$ are across the defect line and the second one is when both fields
are located on the same side with respect the defect line.
With the Wick rotation $it=x$ and the defect line placed at $x=0$,
in the first case we have
\begin{equation}
\begin{array}{c}
<0|\varphi(y,x)\,{\cal{D}}\,\varphi(y_0,-x_0)|0>_E= \\
\\
\begin{displaystyle}
=\int_0^{\infty}\frac{\mbox{d}\theta}{2\pi}e^{-m(x+x_0)\cosh\theta}
e^{im(y-y_0)\sinh\theta}\hat{T}(\theta,c)\, ,
\end{displaystyle}
\end{array}
\end{equation}
whereas in the second case
\begin{equation}
\label{rco}
\begin{array}{c}
\begin{displaystyle}
<0|\varphi(y,x)\varphi(y_0,x_0)\,{\cal{D}}|0>_E=
\int_0^{\infty}\frac{\mbox{d}\theta}{2\pi}\left(e^{-m|x-x_0|\cosh\theta}
e^{im(y-y_0)\sinh\theta}+\right.
\end{displaystyle}
\\
\\
\begin{displaystyle}
\left. +e^{-m(x+x_0)\cosh\theta}
e^{im(y-y_0)\sinh\theta}\hat{R}(\theta,c)\right).
\end{displaystyle}
\end{array}
\end{equation}
The sum of the two contributions
\begin{equation}
\label{gfin}
\begin{array}{c}
G({\bf r},{\bf r}_0;m^2)=
<0|\varphi(y,x)\,{\cal{D}}\,\varphi(y_0,-x_0)|0>_E+ \\
\\
+ <0|\varphi(y,x)\varphi(y_0,x_0)\,{\cal{D}}|0>_E
\end{array}
\end{equation}
is exactly the required Green's function (\ref{gtot}).
Notice that there is another way to compute the same quantity: in fact,
one could mirror the left half plane {\it ab initio} to the right one and
consider the Transmission amplitude as if it was a sort of Reflection
amplitude. One can use this observation in order to define the function
\begin{equation}
K(\theta)=\hat{R}(\theta)+\hat{T}(\theta)
\end{equation}
which together with its crossed counter part defined as
\begin{equation}
\label{KR}
K(\theta)=R\left(\frac{i\pi}{2}-\theta\right)
\end{equation}
satisfy all the conditions (boundary Yang--Baxter, boundary unitarity and
boundary cross--unitarity) a boundary $S$--matrix should fulfill \cite{goza}.
It is thus possible to define the boundary state at the euclidean time
$x=0$
\begin{equation}
\label{bst}
|B>=\exp\left[\frac{1}{2}\int_{-\infty}^{+\infty}
\frac{\mbox{d}\theta}{2\pi}K(\theta)A^{\dagger}(-\theta)
A^{\dagger}(\theta)\right]|0>
\end{equation}
by which we may rewrite our Green's function as
\begin{equation}
\label{last}
G({\bf r},{\bf r}_0;m^2)=
<0|\varphi(y,x)\varphi(y_0,x_0)|B>_E\, .
\end{equation}
\section{Concluding remarks}
Although the original problem of the Random Walk near the compensation point
is defined in half-space, it has been convenient to formulate the dynamics in
terms
of two copies defined for $x>0$ and $x<0$ with appropriate
boundary conditions. In particular, it has been possible to identify the
Transmission amplitude of the defect line model with the compensative role of
the potential in the Random Walk problem and the Reflection amplitude of the
defect line with the {\it hard--wall} limit. For a generic temperature,
the dynamics is ruled by an overlapping of the two contributions, as shown
in equation (\ref{gfin}).
Note that the Quantum Field Theory with action (\ref{action})
could have been solved directly by using the equations for the boundary
$S$--matrix found by Ghoshal and Zamolodchikov \cite{goza}.
Indeed, once the
expansion of the field $\varphi$ for $x\geq0$
\begin{equation}
\varphi(x,t)=\frac{1}{\sqrt{2}}\int\frac{\mbox{d}\theta}{2\pi}
\left[A(\theta)e^{-im(t\cosh\theta-x\sinh\theta)}+
A^\dagger(\theta)e^{im(t\cosh\theta-x\sinh\theta)}\right]
\end{equation}
is inserted into the boundary condition
\begin{equation}
\partial_x\varphi|_{x=0}=c\,\varphi|_{x=0}\, ,
\end{equation}
we have the equation
\[
{\cal{B}}\,A^\dagger(-\theta)=R(\theta){\cal{B}}\,A^{\dagger}(\theta)\, ,
\]
with
\begin{equation}
R(\theta)=\frac{\sinh\theta-ic/m}{\sinh\theta+ic/m}
\end{equation}
and ${\cal{B}}$ the boundary operator.
The Reflection amplitude $R(\theta)$ can be used to define the boundary state
(\ref{bst}) and the Green's function (\ref{last}). However, by using
this approach, the different role played by the Transmission and Reflection
amplitudes of the quantum defect line would have been missed.
As our last remark, it is worth to mention that the action (\ref{action})
has been extensively studied in the context of phase transitions
near surfaces in \cite{BM,surf}, where it describes the {\it high--temperature}
Landau--Ginzburg lagrangian for a magnetic system with boundary. Its
validity is restricted by the occurrence of a surface phase transition for
$c<0$: high--temperature then means temperature higher than the surface
critical temperature, which is $\tau_s=0$ (i.e. $T_s=T_c$) if $c>0$ and
$\tau_s=|c|^2$ if $c$ is negative. Notice that these limits are those
mentioned
after equation (\ref{F}) and were also discussed in the context of boundary
(or defect) scattering amplitudes in \cite{dms}. In the last context,
the poles of $K(\theta)$ for $\tau \leq \tau_s$ simply imply spontaneous
emission
of pairs of particles from the boundary, a condition that destroys the
stability of the system.
\subsection*{Acknowledgments}
The author would like to thank G. Mussardo for the many useful discussions
and helps.
|
\subsection{Near horizon scales: (300 - 3000) ${\rm h}^{-1}$M{\rm pc}}
These scales are so large that the best way to probe them is by
studying the MBR anisotropy at angular scales which correspond to
these linear scales. Since a scale $L$ subtends an angle $\theta(L)
\cong 1^{\circ} (L/100h^{-1}$ Mpc) at $z \simeq
z_D$, the $(\triangle T / T)$ observations at $(3^{\circ} -
30^{\circ})$ probe these scales. The COBE-DMR
observations{\cite{SMOOT92}} of $(\triangle T/T)_{\rm rms}$ and
$(\triangle T / T)_Q$ allow one to obtain the following conclusions:
(i) $\sigma (10^3 h^{-1}$ Mpc)$ \simeq 5 \times 10^{-4}$ (ii) The
power spectrum at large scales is consistent with $P_{\rm in}(k)
\simeq Ak$ and, if we take $\Omega = 1, $, then $A^{1/4} \cong (24 \pm
4)h^{-1}$ Mpc (iii) In this range, $\sigma(R) \cong (24 \pm 4h^{-1}$
Mpc$/R)^2$.
\subsection{ Very large scales : $(80 - 300) h^{-1}$ Mpc}
\subsubsection {CMBR probes:}
These scales span $(0.8^{\circ} - 3^{\circ})$ in the sky at $z \simeq
z_D$. Several ground based and balloon-borne experiments to detect
anisotropy in MBR probe this scale. For example, the UCSB South Pole
experiment has reported{\cite{SCHUSTJ93}} a preliminary `detection'
of $(\triangle T/T)\simeq 10^{-5}$ at $1.5^{\circ} $ scale, and a 95\%
confidence level bound of $(\triangle T/T) < 5 \times 10^{-5}$. This
translates into the constraint of $\sigma (10^2h^{-1}$ Mpc) $\la 5
\times 10^{-2}$.
The angular anisotropy of CMBR is dominated by the gravitational
potential wells of dark matter at large scales. However, at $\theta
\simeq 1^{\circ}$ baryonic process affect the pattern of anisotropy
significantly. The precise determination of degree scale anisotropy
can, therefore, help in distinguishing between different
models{\cite{WHITESCOTT94}}.
\subsubsection{ Galaxy surveys:} Some galaxy surveys, notably CfA2
survey and pencil-beam surveys probe scales which are about
$10^2h^{-1}$ Mpc in depth{\cite{BROAD90}}. Unfortunately, the
statistics at these large scales is not good enough for one to obtain
$\sigma(R) $ directly from these surveys.
\subsection{ Large scales : ~(40 - 80)~ $h^{-1}$~ Mpc}
\subsubsection{ CMBR probes:}
The scales correspond to $\theta_{MBR} \simeq (24' - 48')$ and are
probed by the experiments looking for small angle anisotropies in MBR.
The claimed detection{\cite{CHENG93}} by MIT-MASM of $(\triangle T/T)
\cong (0.5 - 1.9) \times 10^{-5}$ at $\theta \simeq 28', $ if
confirmed, will give a bound of $\sigma (50h^{-1}$ Mpc $) \la 0.3.$
\subsubsection{ Galaxy Surveys:}
Several galaxy surveys, in particular the IRAS-QDOT and APM surveys,
give valuable information about this range{\cite{RR90}}. The angular
correlation of galaxies, measured by APM survey is $\omega (\theta)
\simeq (1-5) \times 10 ^{-3} $ at $\theta \simeq 14^{\circ} $. This
corresponds to $\sigma (50 h^{-1}$ Mpc$)\cong 0.2$. What is more
important, these surveys can provide valuable information about the
shape of the power spectrum in this range if we assume that galaxies
faithfully trace the underlying mass distribution.
\subsubsection{Large scale velocity field:}
Using distance indicators which are independent of Hubble constant, it
is possible to determine the peculiar velocity field $v(R)$ of
galaxies upto about $80h^{-1}$ Mpc or so. The motion of these galaxies
can be used to map the underlying gravitational potential at these
scales. Careful analysis of observational data shows{\cite{DEKEL94}}
that $v(40h^{-1}$ Mpc$)\simeq (388 \pm 67)$ kms$^{-1}$ and $v
(60h^{-1}$ Mpc$) \simeq (327 \pm 82)$ kms$ ^{-1}$. From these values
it is possible to deduce that $\sigma (50h^{-1}$ Mpc $) \simeq 0.2$.
These observations also allow us to determine the value of the
parameter $(\Omega^{0.6}/b_{\rm IRAS})$ where $b_{\rm IRAS}$ is the
bias factor with respect to IRAS galaxies. One finds that
$(\Omega^{0.6} / b_{\rm IRAS}) = 1.28^{+ 0.75}_{-0.59}$ which implies
that if $\Omega = 1$, then $b_{\rm IRAS} = 0.78 ^{+0.66}_{-0.29}$ and
if $b_{\rm IRAS} = 1$ then $\Omega = 1.51 ^{+1.74}_{-0.97}.$
\subsubsection{Clusters and voids:} The cluster-cluster corelation
function and the spectrum of voids in the universe can, in principle,
tell us something about these scales. Unfortunately, the observational
uncertainties are so large that one cannot yet make quantitative
predictions.
\subsection{Intermediate scales : $(8 - 40) h^{-1}$ Mpc}
\subsubsection{ Galaxy Surveys:} The galaxy - galaxy correlation
function $\xi_{\rm gg} \cong [r / 5.4 h^{-1}Mpc]^{-1.8}$ is fairly
well determined at these scales. Direct observations suggest that
$\sigma_{\rm gal} (8h^{-1}$ Mpc$ )\simeq 1$ but the $\sigma_{\rm DM}$
and $\sigma_{\rm gal}$ at these scales can be quite different because
of possible biasing.
\subsubsection{ Cluster Surveys:} There have been several attempts to
determine the correlation function of clusters of different classes.
It is generally believed that $\xi_{\rm cc} \simeq (r/L)^{-1.8}$ with
$L\simeq 25 h^{-1}$ Mpc. The index $n = 1.8$ is fairly well determined
though the scale $L$ is not; in fact, $L$ seems to depend on the
richness class of the cluster. The quantity $(\xi_{\rm cc}/\xi_{\rm
gg})^{1/2}$ can be thought of as measure of the relative bias between
cluster and galaxy scales. Observations suggest{\cite{DALTON92}} that
this quantity depends on the cluster class and varies in the range $(2
- 8)$. The observational uncertainties are still quite large for this
quantity to be of real use; but if the observations improve we will
have valuable information from $\xi_{cc}$.
\subsubsection{ Abundance of rich clusters:} The scale $R=8h^{-1}$ Mpc
contain a mass of $1.2 \times 10^{13} \Omega h^{-1}_{50} M_{\odot}$.
When this scale becomes nonlinear, it will reach an overdensity of
about $\delta \simeq 178$, or -- equivalently -- it will contract to a
radius of $R_f \simeq (8h^{-1}$ Mpc) $/(178)^{1/3} \simeq 1.5 h^{-1}$
Mpc. A mass of $10^{15}M_{\odot}$ in a radius of $1.5 $ Mpc is a good
representation of Abell clusters we see in the universe. {\it This
implies that the observed abundance of Abell clusters can be directly
related to $\sigma (8 h^{-1}$} Mpc). Several people have attempted to
do this{\cite{SDMWHITE93}}; the final results vary depending on the
modelling of Abell clusters, and give $\sigma (8h^{-1}$ Mpc$) \simeq
(0.5 - 0.7)$. Since $\sigma_{\rm gal}(8h^{-1}$ Mpc)$\simeq 1$, this
shows that $b \simeq (1.23 - 2) $ at $ 8h^{-1}$ Mpc.
It is possible to give this argument in a more general
context{\cite{KSTP94}}. Suppose that the contribution to critical
density from collapsed structures with mass larger than $M$ is
$\Omega(M)$, at a given redshift $z$. Then one can show that
$$ \Omega (M) = erfc \left[ { \delta_c (1 + z) \over \sqrt 2
\sigma_0(M) } \right]$$
where $\delta_c = 1.68$ and erfc(x) is the complementary error
function. The Abell clusters (at $z = 0$) contribute in the range
$\Omega \simeq (0.001 - 0.02) $. Even with such a wide uncertainty,
we get $\sigma_{\rm clus} \simeq (0.5 - 0.7) $.
\subsection{Small scales : $(0.05 - 8) h^{-1}$ Mpc}
These scales correspond to structures with $M_{\rm smooth} \simeq (3
\times 10^8 - 1.2 \times 10^{15}) ~\Omega h^{-1}_{50} M_{\odot}$ and
we have considerable amount of observational data covering these
scales. Unfortunately, it is not easy to make theoretical predictions
at these scales because of nonlinear, gas dynamical, effects.
\subsubsection{ Epoch of galaxy formation:} Observations indicate that
galaxy-like structures have existed even at $z \simeq 3$. This
suggests that there must have been sufficient power at small scales to
initiate galaxy formation at these high redshifts. Unfortunately, we
do not have reliable estimate for the abundance of these objects at
these redshifts and hence we cannot directly use it to constrain
$\sigma (R)$.
\subsubsection{ Abundance of quasars:} The luminosity function of
quasars is fairly well determined upto $z \approx 4$. If the
astrophysical processes leading to quasar formation are known, then
the luminosity function can be used to estimate the abundance of host
objects at these redshifts. Though these processes are somewhat
uncertain, most of the models for quasar formation suggest that we
must have $\sigma (0.5h^{-1}$ Mpc$) \ga 3.$
\subsubsection{Absorption systems:} The universe at $1 \la z \la 5$ is
also probed by the absorption of quasar light by intervening objects.
These observations suggest that there exist significant amounts of
clumped material in the universe at these redshifts with neutral
hydrogen column densities of $N_{\rm HI} \simeq (10^{15} -
10^{22}) $cm$^{-2}.$ We can convert these numbers into abundances of
dark matter halos by making some assumptions about this structure. We
find that{\cite{KSTP94}} in the redshift range of $z \simeq (1.7 -
3.5)$ damped Lyman alpha systems contribute a fractional density of
$\Omega_{Ly} \simeq (0.06 - 0.23).$ This would require
$\sigma(10^{12}M_{\odot}) \simeq (3 - 4.5)$.
\subsubsection{ Gunn-Peterson bound:} While we do see absorption due
to {\it clumped } neutral hydrogen, quasar spectra do not show any
absorption due to smoothly distributed neutral hydrogen. Since the
universe became neutral at $z \la z_D \simeq 10^3$, and since galaxy
formation could not have made all the neutral hydrogen into clumps, we
expect the IGM to have been ionised sometime during $5 \la z \la
10^3$. It is not clear what is the source for these ionising
photons. Several possible scenarios (quasars, massive primordial
stars, decaying particles etc.) have been suggested in the literature
though none of these appears to be completely satisfactory. In all
these scenarios, it is necessary to form structures at $z \ga 5$ so
that an ionising flux of about $J = 10^{-21} $ergs cm$^{-2}$s$^{-1}$
Hz$^{-1}$ sr$^{-1}$ can be generated at these epochs. Once again, it
is difficult to convert this constraint into a firm bound on $\sigma$
though it seems that $\sigma (0.5 h^{-1}$ Mpc$) \ga 3 $ will be
necessary.
\section {Gravitational lensing and large scale structure}
In the above discussion we have not taken into consideration the
constraints imposed by gravitational lensing effects on the structure
formation models. This aspect will be discussed in detail in the other
articles in this volume; here we shall contend ourselves with a brief
mention of the possibilities.
Gravitational lensing probes the gravitational potential directly and
can provide valuable information at very different scales. At the
largest scales $(R \simeq 10^3 Mpc)$ lensing can be used to probe the
geometry of the universe. For example, it is possible to put firm
bounds on the energy contributed by cosmological constant from such
considerations.
At intermediate scales ( $R \simeq 50 Mpc)$ lensing has the potential
of providing information about the power spectrum of fluctuations
which are in the quasilinear phase. In principle the distortion of
images can be inverted to obtain this information, though in practice
this is extremely difficult.
At smaller scales, the ``weak lensing'' -- leading to arcs and
arclets at cluster scales -- is already providing a clue to the
mapping of dark matter distribution in clusters. On the other hand,
direct optical and X-ray observations provide us information about
the distribution of visible matter in clusters. The combination of
these techniques should give us valuable information as regards the
dynamical processes which separated baryons from dark matter.
At still smaller scales, galactic potentials have the capacity to
produce multiple images of distant sources. The statistics of these
multiple images depends crucially on the core radii of the galaxies,
which in turn depends sensitively on the structure formation models.
The absence of significant number of multiple images with large
angular separations puts severe constraints on models for structure
formation. The analytic modelling of nonlinear dark matter clustering
described earlier could be used to strengthen these constraints still
further.
\section{ Scorecard for the models}
The simplest models one can construct will contain a single component
of dark matter, either cold or hot. Such models are ruled out by the
observations. The HDM models, normalised to COBE result will have
maximum power of $\triangle_m \cong 0.42h^{-2} (m/30eV)^2 $ at $k =
k_m = 0.11$ Mpc$^{-1}(m/30eV)$. In such a case, structures could have
started forming only around $(1 + z_c) \cong (\triangle_m / 1.68)
\cong h^{-2}_{50} (m/30eV)^2$ or at $z_c \cong 0 .$ We cannot explain
a host of high-$z$ phenomena with these models. The pure CDM models
face a different difficulty. These models, normalised to COBE, predict
$\sigma_8 \simeq 1$, which is too high compared to the bounds from
cluster abundance. When nonlinear effects are taken into account, one
obtains $\xi_{\rm gg} \propto r^{-2.2}$ for $h = 0.5$ which is too
steep compared to the observed value of $\xi_{\rm gg} \propto
r^{-1.8}$. In other words, CDM models have wrong shape for $\xi(r)$
to account for the observations.
The comparison of CDM spectrum with observations suggests that we
need more power at large scales and less power at small scales. This
is precisely what happens in models with both hot and cold dark matter
or in models with nonzero cosmological constant. These models have
been extensively studied during the last few years, and they fare
well as far as large and intermediate scale observations are
concerned. However, they have considerably less power at small scales
compared to CDM model. As a result, they do face
difficulties{\cite{KSTP94}} in explaining the existence of high
redshift objects like quasars. For example, a model with $30\%$ HDM
and $70\%$ CDM will have $\sigma_{0.5} \simeq 1.5$; to explain the
abudnance of quasars comfortably one needs $\sigma_{0.5} \simeq 3.0$.
To explain the abundance of damped Lyman alpha systems one requires
still larger vlaues of about $\sigma_{0.5} \approx 4 $ or so.
Demanding that $\sigma (10^{12}M_{\odot}) > 2$ [which is equivalent
to saying that $10^{12}M_{\odot}$ objects must have collapsed at a
redshift of $z_{12} = (2 / 1.68) -1 \simeq 0.2]$ will completely rule
out this model. Similar difficulties exist in models with cosmological
constant. Notice that all models are normalised using COBE results at
very large scales. Hence the severest constraints are provided by
observations at smallest scales, since the ``lever-arm'' is longest in
that case.
The comparison of models show that it is not easy to accommodate all
the observtions even by invoking two components to the energy density.
(These models also suffer from serious problems of fine-tuning). By
and large, the half-life of such quick-fix models seem to be about 2-3
years. One is forced to conclude that to make significant progress it
is probably necessary to perform a careful, unprejudiced analysis of:
(a) large scale observational results and possible sources of error
and (b) small scale baryonic astrophysical processes.
|
\section{Introduction\label{in}}
We report here about the activity of the working group
on ``Pion (Kaon) and Sigma Polarizabilities'' we conducted, in collaboration
with R. Baldini, at the Workshop on Chiral Dynamics:
Theory and Experiment, held July 25-29,1994 at MIT. The goal of
this working group was to identify the processes that are more suitable
to measure the electric and magnetic polarizabilities of the abovementioned
hadrons and probe chiral dynamics in the photon-pion (-kaon) and photon-sigma
physics.
The agenda of the group was a mixture of theory and
experiment that allowed us to summarize the current status in this
field, determine what is to be done in order to improve it, from both
the theoretical and the experimental side, and quantify the level of
accuracy needed to make the improvement significant.
We considered the following general areas:
1. Theoretical predictions and models for Compton scattering
$\gamma\pi(K)\rightarrow\gamma\pi (K)$, as well as for
$\gamma\gamma\rightarrow\pi\pi (K K)$,
and the relation to the pion (kaon) polarizabilities.
2. Experiments to measure the pion (kaon) and sigma polarizabilities.
3. Phenomenology required in order to extract the polarizabilities
from the experimental data.
We have considered the following methods of measurement:
i) Radiative photoproduction of the pion and extrapolating to the pion pole,
in order to extract the polarizabilities form the data.
ii) Experiments to measure the pseudoscalar meson polarizabilities using the
Primakoff effect.
iii) The measurements at the Frascati DA$\Phi$NE with the KLOE detector.
Murray Moinester illustrated the plans at FNAL and the reaction
$\pi\rho\rightarrow\pi\gamma$ in connection with the pion and sigma
polarizabilities.
Thomas Walcher went over the MAMI project at Mainz. Annalisa D'Angelo
discussed the potential of the Graal synchrotron light facility in
Grenoble to carry out polarizability measurements.
S. Kananov discussed the need for a careful estimate of the
radiative corrections for the FNAL experiments where the pion
polarizabilities can be measured.
A presentation of the DA$\Phi$NE capabilities has been done
in the workshop.
The related talks had the following titles:
A. D'Angelo: "The experimental plans in Grenoble",
T. Walcher: "The MAMI experiment at Mainz",
M. Moinester: "Pion polarizabilities and quark-gluon plasma signatures",
S. Kananov: "Radiative corrections for pion polarizability experiments",
In the second part of the working group, devoted to the theoretical
contributions, we had the following talks:
J. Kambor: "Determination of a $O(p^6)$ counterterm from sum rules",
M. Knecht: "$\gamma\gamma\rightarrow\pi^0\pi^0$ and $\pi^0$ polarizabilities
in generalized chiral perturbation theory",
M. Pennington: "Dispersion relations and pion polarizabilities",
S. Bellucci: "Difficulties in extrapolating to the pion pole the data on
radiative pion photoproduction".
\setcounter{equation}{0}
\setcounter{subsection}{0}
\section{Experimental plans for polarizability measurements\label{exp}}
\subsection{Radiative pion photoproduction}
Let us consider first of all the unexpected and very innovative
contribution by Annalisa D'Angelo. She has reported about the
possibility to measure the pion polarizability by Graal, the new
facility at the electron storage ring ESRF in Grenoble.
The Graal facility consists of a tagged and highly polarized
$\gamma$-ray beam, produced by the backscattering of Laser light against the
high energy electrons circulating in the storage ring ESRF at
Grenoble \cite{z1}-\cite{z3}.
If commercial Ion-Argon and Nd-Yag Lasers are used, either linearly or
circularly polarized, a $\gamma$-ray beam of energy ranging from
about 300 MeV to 1.5 GeV with a degree of polarization higher then 70$\%$ over
almost the entire spectrum may be obtained, by appropriate choice of the
Laser line.
A large solid angle multi purpose detector
\cite{z4} is part of the Graal facility
and it will be used to perform experiments on photo-nuclear reactions
\cite{z5}.
It consists of a crystal ball made of 480 BGO crystals (24 cm long)
covering all
azimuthal angles for polar angles in the interval between 25$^{\circ}$ and
155$^{\circ}$. It may be used as electromagnetic calorimeter, with measured
energy resolution of 2$\%$ FWHM for 1 GeV photons, or to detect protons of
energy
up to 300 MeV \cite{z6}-\cite{z8}.
The central hole of the BGO ball ($\phi$ = 20 cm) will contain a barrel of 32
plastic scintillators; it will be used to discriminate between charged and
neutral particles and to identify the charged particles with the
$\frac{\Delta E}{E}$
technique. Inside the barrel two
cylindrical wire chambers will provide
improved angular resolution for the reconstruction of the trajectories of the
charged particles.
In the forward direction two detectors, each consisting of three plane wire
chambers rotated of 45$^{\circ}$, and a scintillating wall will provide charged
particle tracking information and TOF measurements; 10$\%$ efficiency is
expected for neutron detection in the scintillating wall.
We have started to investigate the possibility of using the Graal facility to
study the radiative pion photoproduction from the proton (namely the reaction
$\vec{\gamma} + p \rightarrow n + \pi^{+} + \gamma$), in order to extrapolate
the experimental data to the pion pole and determine the cross-section of the
Compton scattering on the pion \cite{z9}.
The interest of this measure, in order to get information on the
pion polarizability, has been pointed out, among other authors, by D.
Drechsel and L.V. Fil'kov \cite{z9};
they have stressed that in order to obtain a
reliable extrapolation it is necessary to have experimental data in kinematical
condition as close as possible to the point $t=0$, being $t$ the momentum
transferred between the final neutron and the initial proton.
A measurement performed at 1.5 GeV photon
energy, using polarized photons, would
fulfill the experimental requirements of Ref. \cite{z9}, also providing higher
sensitivity to the pion polarizability contributions through the polarization
structure functions.
A fundamental issue of the experimental set-up is the capability
of discriminating the reaction of interest from the background events, like
those coming from asymmetric decay of $\pi ^{0}$ in the $\pi^+ \pi^0 n$
reaction channel.
In principle all these requirements are fulfilled by the Graal facility:
the scattered photons may be detected in the BGO ball for laboratory angles
between $25^{\circ}$ and $155^{\circ}$; low energy neutrons my be detected in
the forward direction using the plastic wall and they may be identified using
the TOF information; finally the pions may be detected at all
angles by the wire
chambers with good angular resolution. All these experimental information
should allow a complete reconstruction of the interesting events in selected
kinematical conditions, with expected good background rejection.
The Graal facility set-up is therefore a promising tool to perform the
first experiment with polarized photons on radiative pion
photoproduction in order to extract information on the pion
polarizability.
\\
Thomas Walcher has reviewed, from his general talk on
the experimental activity at MAMI in Mainz,
the measurement of the pion polarizability still by means of radiative pion
photoproduction.
Walcher has shown some kinematical conditions suitable
for the measurement of the charged pion polarizability.
The sensitivity in the Chew-Low extrapolation at the pion pole has been
stressed \cite{z9}. For instance a variation of ${\alpha}_{\pi}$
from 0 up to 7 is equivalent to a 20 $\%$ variation in the extrapolated
amplitude. Hence the extrapolation has to be done at the
2 $\%$ level , if $\alpha_{\pi}$ has to be measured at a 10 $\%$
level. This sensitivity depends, of course,
on the minimum momentum transfer
$t_{min}$ achieved.
For instance for a 700 MeV incident photon,
a 152 MeV final photon, a neutron in the angular range
$12^{0}$-$32^{0}$ and in the energy range 400 - 350 MeV, it is
$t$= 0.31, in pion mass squared units. Nucleon, pion and $\Delta(1236)$
pole diagrams have been evaluated and the extrapolation seems
feasible.
Conversely the measurement of the neutral pion polarizability is not realistic
at the MAMI energies. The situation may be improved at higher
incident photon energies and a polarized beam would be very welcome,
just like the Graal facility!
By the way M. Moinester has stressed that the maximum energy
meaningful for extracting the pion polarizability is about 2 GeV,
corresponding
to the $\rho$ mass in the photon-pion c.m. system.
The MAMI detector consists of a MWPC system to detect the charged
pion, close to the incident
photon beam, a segmented $BaF_{2}$ to detect the scattered photon
and a system of scintillators for detecting the neutron and
providing the neutron time of flight.The expected yield
of radiative pion photoproduction is
$\simeq$ 2000 events/day. The background due to double
photoproduction, simulating radiative pion photoproduction,
is $\simeq$ 160 events/day.
\subsection{Primakoff effect}
The contribution of Murray Moinester has concerned the measurement of
$\pi$ ($K$) and $\Sigma$ polarizabilities via $\pi$ ($K$) and $\Sigma$
high energy beams at Fermilab by E781 \cite{z10}.
This experimental activity has been reported already
in detail by M. Moinester in his contribution to the workshop.
Therefore only the main topics are emphasized here, first of all
the relationship between polarizability and radiative transitions.
The $A_{1}$ radiative width is a good illustration of this
statement. It has been demostrated \cite{z11}
according to the current algebra the main
contribution to the pion electric polarizability comes from the exchange
of the $A_{1}$. Xiong, Shuryak and Brown
\cite{z12} have
shown that a radiative width $\Gamma$($A_{1}\rightarrow\pi\gamma$)=
1.4 MeV is needed to get the current algebra value, on the
contrary the experimental value is $\Gamma$($A_{1}\rightarrow\pi\gamma$)=
0.64 $\pm$ 0.25 MeV \cite{z13}. More data from E781 are welcome to
settle this relevant problem.
By the way the unexpected correlation between the reaction
$\pi \rho \rightarrow \pi \gamma$ and the photon flux in a quark-gluon
plasma has been pointed out.
Experimental results from the previous FNAL experiment E272 have
been shown to demonstrate the experimental feasibility of the
Primakoff effect, even if E272 did not get enough statistics
to measure $\alpha_{\pi}$. The main experimental problem in
measuring the Primakoff effect by the new FNAL experiment, E781,
concerns a suitable fast trigger. It is not implemented at the
moment, taking into account the high rate and the difficulties
for using any signal from the scattered photon detector, which
is 50 meters downstream the target.
{}From a theoretical point of view an open question remains the
fair disagreement between the charged $\rho$ radiative width,
as obtained via the Primakoff effect, and the neutral $\rho$
radiative width.
\\
Finally M. Moinester has stressed the role of radiative transitions
in the case of a $\Sigma$ beam, which should also be available in E781.
The radiative transitions to the Sigma*(1385) provide a measure of the s-quark
magnetic moment of the Sigma\cite{lipkin}.
Positive and negative $\Sigma$ are expected to have very
different polarizabilities. In particular the $\Sigma^{-}$ magnetic
moment should
be negligible, both taking into account the 3 quarks have the
same charge and according to the U-spin symmetry. Furthermore
it has been
shown that the event rate for measuring $\Sigma$ radiative
transitions by a 600 GeV $\Sigma$ beam is higher than the
expected rate for measuring the pion Primakoff effect by
E781.
\\
S. Kananov has reported about radiative corrections in the
scattering of pions by nuclei at high energies \cite{z14}. It has been
shown that radiative corrections can simulate a variation of
the magnetic
polarizability $\beta_{\pi}\simeq$ -0.2 from $\beta_{\pi}\simeq$ -5,
with plausible cuts for the outgoing photon.
\subsection{$\gamma\gamma \rightarrow \pi\pi$ at threshold}
Another way to get the pion polarizability is by means of
the measurement of $\gamma\gamma\rightarrow\pi\pi$ at threshold,
performed at the new Frascati $\Phi$-factory DA$\Phi$NE \cite{z15}.
A presentation of this new experimental facility has been
done already in the workshop and no further discussion on experimental
details has been done in this working group.
In summary the new Frascati $e^{+}e^{-}$ storage ring DA$\Phi$NE
is supposed to deliver a luminosity $\simeq 5 \cdot 10^{32}$ cm$^{2}$
$s^{-1}$
at the $\Phi$ mass, with the possibility to increase the total
energy up to 1.5 GeV. Two detectors are under construction:
an all purposes detector, KLOE \cite{z16} mainly dedicated to CP
violation in $K$ decay, and FINUDA \cite{z17} mainly dedicated to hypernuclei
physics.
KLOE is expected \cite{z18} to detect $10^{3} \div 10^{4}$
times the events collected at present in
$e^{+}e^{-} \rightarrow e^{+}e^{-} \pi^{+}\pi^{-}$
and $e^{+}e^{-} \rightarrow e^{+}e^{-} \pi^{0}\pi^{0}$
at threshold.
Unfortunately at the $\Phi$ energy the decay $\Phi\rightarrow K_{S} K_{L}$
is an overwhelming
background (in $\sim$ 15 $\%$ of the events $K_{L}$ are not detected)
and tagging the outgoing $e^{+} e^{-}$ is needed.
Two kind of tagging system for the outgoing leptons are foreseen.
First of all there are two different rings for electrons and positrons
and the splitter magnet after the interaction region is a
suitable magnetic analyzer for the outgoing $e^{+} e^{-}$,
mostly forward emitted. Furthermore $e^{+} e^{-}$
emitted at larger angles are, in part,
detected by the central tracking detector in KLOE.
The $e^{+} e^{-}$ angular distribution depends on $m_{e}/E_{e}$ :
therefore there are more events at large angles in DA$\Phi$NE
respect to the high energy $e^{+} e^{-}$ storage rings.
By the way correlations in the azimuthal angles between the pions and the
outgoing leptons could be performed \cite{z19}, increasing the possibility
to disentangle the D wave contribution. Otherwise $\gamma\gamma$
interactions near threshold are supposed to provide mainly the $\pi\pi$ S wave,
which provides only $\alpha_{\pi} - \beta_{\pi}$.
The overall double tagging efficiency is $\sim$ 15 $\%$ \cite{z18}.
The background from beam-beam bremmstrhalung is still under
study for evaluating the single tagging efficiency.
In $\gamma\gamma$ interactions complications
related to any nuclear target are avoided, but it has been
demonstrated in the following theoretical discussion that the
extrapolation to the pion pole is much more difficult.
Therefore the conclusion has been achieved that
$\gamma\gamma$ interactions are not the best way
to get the pion polarizability. Nevertheless $\gamma\gamma$
interactions near threshold remain a very clean test of any theoretical
description of strong interactions at low energies.
\\
Another possibility pointed out for measuring neutral and
charged $\alpha_{\pi}$ in
$e^{+}e^{-}$ is by means of $e^{+}e^{-} \to \pi\pi\gamma$
increased by the interference both with $\omega \to$ $\pi^{0}$ $\rho^{0}$ $\to$
$\pi_{0}$$\pi^{0}$$\gamma$ and, for the charged one,
also with radiative $\rho$ production.
\setcounter{subsection}{0}
\section{Theoretical issues in polarizability experiments\label{th}}
We begin with the process $\gamma\gamma\rightarrow\pi^0\pi^0$.
In this case the Born amplitude vanishes and the one-loop corrections
in Chiral Perturbation Theory (CHPT)\cite{wein79}-\cite{review} are finite
\cite{bico,dhlin}. The corresponding cross section
is independent of the free parameters of the chiral lagrangian and
does not agree with the experimental measurements at
Crystal Ball \cite{cball}, as well as with
calculations based on dispersion relations \cite{goble}-\cite{kaloshin},
even at low-energy.
The low-energy amplitude recently calculated to two-loops in CHPT
\cite{bgs} agrees with the Crystal Ball data and compares very well with the
results of a dispersive calculation by Donoghue and Holstein \cite{dohod}.
The value of the low-energy constants can be obtained in several ways, e.g.
by resonance exchange. The resonance saturation method provides empirical
values for the
scale-dependent renormalized constants of CHPT \cite{glan}-\cite{glnp}
\begin{equation}
L_i^r (\mu)=L_i^r ({\mu}_0)-\frac{{\Gamma}_i}{16\pi^2} ln\frac{\mu}{{\mu}_0}
{}~\; \; , ~i=1,..,10
\end{equation}
with a scale $\mu$ in the range 0.5 $GeV$ -- 1 $GeV$ and a set of constants
$\Gamma_i$ defined in \cite{glan}-\cite{glnp}.
This method has been used in Ref. \cite{bgs} to pin down the couplings in the
$\gamma\gamma\rightarrow\pi^0\pi^0$ amplitude to order $p^6$.
In his talk J. Kambor discussed how to determine
these couplings from sum rules, exploiting the low- and high-energy behaviour.
Let us consider the vector-vector two-point function
\begin{equation}
i\int d^4x e^{iqx}<0|T(V_{\mu}^a (x)V_{\nu}^b (0)|0> =
\delta^{ab} (q_{\mu}q_{\nu}-g_{\mu\nu}q^2 )\Pi^a (q^2) .
\end{equation}
Following \cite{wein79}-\cite{glnp} and \cite{dohod} we write a
dispersion relation for $\Pi^a (q^2)$
\begin{equation}
\Pi^a (q^2)=\frac{1}{\pi}\int ds\frac{Im\Pi^a (s)}{s-q^2-i\epsilon}
{}~+~subtractions \; \; ,
\label{na0}
\end{equation}
in order to
make contact with the high-energy behaviour of the theory
desumed from the perturbative $QCD$ sum rules
\begin{eqnarray}
\rho_V^a (s)& = &\frac{1}{\pi}Im\Pi^a (s)\; \; ,
\nonumber \\
lim_{s\rightarrow\infty}\rho_V^a (s)& = &\frac{1}{8\pi^2}+O(\frac{1}{s}) ~,~
a=3,8
\end{eqnarray}
showing that the spectral function $\rho_V^a (s)$ at high energy
goes like a constant plus higher-order terms that are suppressed
at least as $\frac{1}{s}$. Hence, the difference between two spectral
functions goes like $s$ at large $s$, and the integral
\begin{equation}
\int ds\frac{\rho_V^3-\rho_V^8}{s}
\end{equation}
converges. Also the once-subtracted dispersion relation for $\rho_V^3 (s)$
converges
\begin{equation}
\int ds\frac{\rho_V^3}{s^2}~.
\end{equation}
{}From the dispersion relation
\begin{equation}
\Pi^3 (q^2)-\Pi^8 (q^2)=\frac{1}{\pi}\int ds
\frac{Im\Pi^3 (s)-Im\Pi^8 (s)}{s-q^2-i\epsilon}
\label{na1}
\end{equation}
the sum rule is readily obtained
\begin{equation}
\Pi^3 (0)-\Pi^8 (0)=\int ds
\frac{\rho_V^3-\rho_V^8}{s}~.
\label{na2}
\end{equation}
Taking the $q^2$ derivative of the dispersion relations (\ref{na1})
and (\ref{na0}) yields the sum rules
\begin{equation}
\frac{d}{dq^2}(\Pi^3 (0)-\Pi^8 (0))=\int ds
\frac{\rho_V^3-\rho_V^8}{s^2}
\label{na3}
\end{equation}
and, respectively,
\begin{equation}
\frac{d}{dq^2}\Pi^3 (0)=\int ds
\frac{\rho_V^3}{s^2}~.
\label{na4}
\end{equation}
Notice that if one takes too many $q^2$-derivatives, then the integrals
become dominated by the threshold region.
As for the low-energy behaviour, J. Kambor showed how to use CHPT to calculate
$\Pi^a (q^2)$ for small $q^2$ values. This calculation is carried out in
$SU(3)\times SU(3)$ to the two-loop order, and the result depends on
the $O(p^4)$ and $O(p^6)$ low-energy constants $L_i$ and, respectively,
$d_j$ \cite{GK}. The integral on the r.h.s. of Eq. (\ref{na2}) can be
evaluated from the $e^+ e^-$ data.
In the narrow width approximation
one gets the following estimate \cite{KMS}:
\begin{equation}
\int ds\frac{\rho_V^3-\rho_V^8}{s}=\frac{3}{4\pi\alpha^2}
\biggr( \frac{\Gamma_{\rho\rightarrow e^+ e^-}}{M_{\rho}}
-3\frac{\Gamma_{\omega\rightarrow e^+ e^-}}{M_{\omega}}
-3\frac{\Gamma_{\phi\rightarrow e^+ e^-}}{M_{\phi}}\biggr)
=(11.1\pm 2.0)\cdot 10^{-3} ~.
\end{equation}
Thus, the integration region becomes
divided into three pieces, i.e. $4M_{\pi}^2\le s\le \Lambda_1$,
$\Lambda_1\le s\le\Lambda_2$, and $s\ge\Lambda_2$. Here we denote
by $\Lambda_{1,2}$ two cutoff values of about 0.4 $GeV$ and 2 $GeV$,
respectively, and $M_{\pi}$ is the pion mass. In the first region
the shape of the spectral function is obtained from the two-loop CHPT
calculation, i.e. $\rho_V =\rho_V^{1-loop}+\rho_V^{2-loop}$. In
the second (third) region the $e^+ e^-$ data (the perturbative $QCD$
calculation) can be used to obtain $\rho_V (s)$.
The result of this calculation (once it is completed) yields a scale
independent determination of $d_3$ contributing to
$\gamma\gamma\rightarrow\pi^0\pi^0$. J. Kambor expects an accuracy for
this estimate between 10 and 20 percent. This would be more precise than
the estimate carried out in \cite{KMS}
using a similar procedure (excluding, however, the two-loop contribution),
within the framework of the Generalized Chiral Perturbation Theory (GCHPT)
\begin{equation}
d_3 = (9.4\pm 4.7)\cdot 10^{-6} ~.
\label{na5}
\end{equation}
There are two more sum rules to consider. In particular Eq. (\ref{na3})
is effectively a sum rule for $L_9$ (the $d_i$ contributions to
$\Pi^3$ and $\Pi^8$ drop out of the sum rule), whereas Eq. (\ref{na4})
is a sum rule for $d_5$, $d_6$ contributing to
$\gamma\gamma\rightarrow\pi^+\pi^-$ and
$\gamma\pi^{\pm}\rightarrow\gamma\pi^{\pm}$. A similar treatment can be
applied to the sum rule for the axial-axial two-point function. This
gives an estimate of the low-energy constants $L_{1,2,3}$ contributing
to $\pi\pi$-scattering.
The GCHPT approach is described in \cite{reffks} (see also M. Knecht's
talk in the $\pi\pi$-scattering Working Group Section of these Proceedings).
Within this approach the cross section for
$\gamma\gamma\rightarrow\pi^0\pi^0$ and the pion polarizabilities
have been calculated up to and including $O(p^5 )$
\cite{KMS}. M. Knecht discussed the result of this calculation.
He showed that the cross section depends on the quark
mass ratio $r=\frac{2m_s}{m_u +m_d}$ and is consistent with the data
from Crystal Ball \cite{cball}, provided the value of the ratio is
at least a factor of two or three smaller than its standard value
in CHPT \cite{KMS}. M. Knecht showed that a
low-energy theorem analogous to the one
outlined above, but without taking into account the 2-loop contribution,
yields, through the evaluation of the dispersive integral via resonance
saturation, the following value for the $O(p^5 )$ constant $c$
defined in \cite{KMS}:
\begin{equation}
c = -\frac{1}{r-1}(4.6\pm 2.3)\cdot 10^{-3} ~.
\label{na6}
\end{equation}
In the standard CHPT case, i.e. for
\begin{equation}
r = 2\frac{M_K^2}{M_{\pi}^2}-1=25.9 ~,
\label{na7}
\end{equation}
the expression for $c$ given in Eq. (\ref{na6}) corresponds to
the value reported in Eq. (\ref{na5}). This is to be compared with
the value calculated from Appendix D of Ref. \cite{bgs}, using
resonance exchange
\begin{equation}
d_3 = \pm 3.9\cdot 10^{-6} ~.
\end{equation}
M. Pennington reviewed the dispersion relation treatment of the
$\gamma\gamma\rightarrow\pi^0\pi^0$ amplitude that represents
the data quite well \cite{pe}, \cite{pehan}. He showed how to
calculate the cross section from first principles, using a
relativistic and causal description based on the unitarity
of the scattering matrix. The prediction for low-energy
$\gamma\gamma\rightarrow\pi^0\pi^0$ and
$\gamma\gamma\rightarrow\pi^+\pi^-$ is based on the present
knowledge of the $\pi\pi$ phases and $PCAC$. M. Pennington
argued that precision measurement of
$\gamma\gamma\rightarrow\pi^0\pi^0,\pi^+\pi^-$ at the Frascati
$\Phi$-factory DA$\Phi$NE will restrict further the $\pi\pi$ phases
and determine where the chiral zero appears on-shell.
The electric and magnetic polarizabilities enter the
low-energy limit of the coupling
with the photon in the Compton amplitude
for any composite system. The dynamics of hadronic systems
can be probed by measuring the hadron
polarizabilities \cite{revpol}. In particular, the pion Compton scattering
can be investigated in this connection. The charged pion Compton amplitude
\begin{equation}
\gamma(q_1) \pi^+(p_1) \rightarrow \gamma(q_2)
\pi^+(p_2)\; \; ,
\end{equation}
admits an expansion near threshold
\begin{eqnarray}
T^C &=& 2 \left[ \vec{\epsilon}_1 \cdot \vec{\epsilon}_2 \,\! ^\star
\left(
\frac{\alpha}{M_{\pi}} - \bar{\alpha}_{\pi} \omega_1 \omega_2
\right)- \bar{\beta}_{\pi} \left(\vec{q}_1 \times \vec{\epsilon}_1
\right) \cdot \left(
\vec{q}_2 \times \vec{\epsilon}_2 \,\! ^\star \right)
+ \cdots \right]
\end{eqnarray}
with $q_i^\mu = (\omega_i, \vec{q}_i)$.
Below we denote
\begin{eqnarray}
(\alpha \pm \beta)^C &=& \bar{\alpha}_{\pi} \pm \bar{\beta}_{\pi}
\; \; , \nonumber \\
(\alpha \pm \beta)^N &=& \bar{\alpha}_{\pi^0} \pm \bar{\beta}_{\pi^0} \; \; ,
\end{eqnarray}
for charged and neutral pions, respectively.
The charged pion polarizabilities have been determined in an experiment
on the radiative
pion-nucleus scattering $\pi^-A\rightarrow \pi^-\gamma A $ \cite{serpukov1}
and in the pion photoproduction process $\gamma p
\rightarrow \gamma \pi^+n$ \cite{lebedev}.
Assuming the constraint $(\alpha + \beta)^C=0$
the two experiments yield\footnote{Throughout the following, we
express the values of the polarizabilities in units of $10^{-4} fm^3$ }
\begin{eqnarray}
(\alpha - \beta)^C =\left\{ \begin{array}{ll}
13.6 \pm 2.8 & \cite{serpukov1} \\
\; \; \; 40 \pm 24 & \cite{lebedev} \; .
\end{array}
\right.
\label{cs3}
\end{eqnarray}
Relaxing the constraint
$(\alpha + \beta)^C = 0$, one obtains from the Serpukhov data
\begin{eqnarray}
(\alpha + \beta)^C&=& \; \; 1.4 \pm 3.1 (\mbox{stat.}) \pm
2.5 (\mbox{sys.}) \; \; \cite{serpukov2} \; \; , \nonumber \\
(\alpha - \beta)^C&=& 15.6 \pm 6.4 (\mbox{stat.}) \pm 4.4 (\mbox{sys.})
\; \; \cite{serpukov2} \; .
\label{cs4}
\end{eqnarray}
At one-loop in CHPT one has \cite{lopol,z11,bbgm0}
\begin{eqnarray}
\bar{\alpha}_{\pi^0}=-\bar{\beta}_{\pi^0}=
-\frac{\alpha}{96\pi^2M_{\pi}F^2} = -0.50 \; \; .
\label{cs6}
\end{eqnarray}
At order $O(p^6)$ it was calculated in Ref. \cite{bgs}
\begin{eqnarray}
\bar{\alpha}_{\pi^0}&=&-0.35 \pm 0.10\; \; , \nonumber \\
\bar{\beta}_{\pi^0}&=& \; \; \; 1.50 \pm 0.20.
\end{eqnarray}
The low-energy $\gamma \gamma \rightarrow \pi^+\pi^-$ data
\cite{dcharged} have been used in Ref. \cite{bbgm0} to obtain
information on $\bar{\alpha}_\pi$ and $\bar{\beta}_\pi$.
The result in \cite{bbgm0} yields the
numerical value for the leading-order
$\bar{\alpha}_\pi = 2.7 \pm 0.4$, plus systematic uncertainties due to
the $O(p^6 )$ corrections. The latter are not yet available.
A part of the corrections to the charged pion polarizabilities beyond the
one-loop order has been obtained in Refs. \cite{kaloshin86,ba2loop}
including the meson resonance contribution.
M. Knecht analyzed the $O(p^5)$ calculation from Ref. \cite{KMS}
of the $\pi^0$ polarizabilities
in GCHPT. The result $(\alpha +\beta)^N =0$ remains valid to this
order, whereas the remaining combination depends on the quark
mass ratio $r$. Hence this combination can have positive values
for $r$ much less than its standard CHPT value (\ref{na7}),
e.g. $(\alpha -\beta)^N =1.04\pm 0.60$ for $r=10$ \cite{KMS}.
For comparison we recall the standard CHPT prediction
for both combinations to the $O(p^6)$ order \cite{bgs}
\begin{eqnarray}
(\alpha +\beta)^N &=& \; \; \; 1.15\pm 0.30 \; \; , \nonumber \\
(\alpha -\beta)^N &=&-1.90\pm 0.20 .
\end{eqnarray}
The reason for the sign difference of $(\alpha -\beta)^N$ in GCHPT
with respect to the standard CHPT value has been traced back
by M. Knecht to a dominance by the positive $O(p^5)$ contribution
over the strongly suppressed negative
$O(p^4)$ contribution in GCHPT. This suppression is
related to a shift in the
position of the chiral zero, as $r$ becomes much smaller than the
standard CHPT value (\ref{na7}) \cite{KMS}.
Starting from the unitarized $S$-wave amplitudes
for neutral pions, M. Pennington displayed a proportionality
relation between $(\alpha - \beta)^N$ and the position of the chiral zero
$s_N$, showing that the former assumes values between -0.6
and -2.7 as the latter runs from $\frac{1}{2}M_{\pi}^2$ and
$2M_{\pi}^2$. He also discussed the validity of the errors
quoted in a recent estimate of
$(\alpha + \beta)^{C,N}$ by Kaloshin and collaborators \cite{kalo}.
Here the polarizabilities appear as adjustable parameters
in the unitarized $D$-wave amplitudes, hence the values
of $(\alpha + \beta)^{C,N}$ can be determined from the data
with the result \cite{kalo}
\begin{eqnarray}
(\alpha + \beta)^C &=& 0.22 \pm 0.06\; \; \cite{dcharged} \; \; , \nonumber \\
(\alpha + \beta)^N&=& 1.00 \pm 0.05 \; \; \cite{cball} \; .
\label{cs5}
\end{eqnarray}
M. Pennington, arguing on the partial wave analysis
of the data that shows large uncertainties even
at the $f_2$(1270) mass, concluded that the errors quoted in
(\ref{cs5}) for $(\alpha + \beta)^N$ are unbelievably small. His final conclusion,
that one must measure $\gamma\pi\rightarrow\gamma\pi$ can be
wholeheartedly shared, in view of future measurements of the
pion polarizabilities. In this respect, it is very important to devise
fully reliable methods that allow to extract the pion Compton scattering
amplitude from the measurement of the radiative pion photoproduction,
as discussed by the author.
\vspace{1.5cm}
\noindent
{\bf Acknowledgements}
It is a pleasure to thank the organizers of this Workshop for
a very stimulating working environment and the participants
to this working group for their precious help in preparing this report.
Very valuable help from R. Baldini in the preparation of this
work is also gratefully acknowledged.
\newpage
|
\section{Introduction}}
\newcommand{\section{Conclusions}}{\section{Conclusions}}
\newcommand{\section*{Acknowledgments}}{\section*{Acknowledgments}}
\begin{document}
\hyphenation{transfor-ma-tion introdu-ce Lo-ren-tz qua-ter-ni-on
transformati-ons manipulati-ons com-ple-te intro-du-ce com-plexi-fi-ed
Di-rac in-ter-national}
\begin{titlepage}
\begin{center}
August, 1995 \hfill {\footnotesize hep-th/95mmnnn}
\vs{3cm}
{\Large \bf Quaternions and Special Relativity}
\vs{2cm}
{\sc Stefano De Leo}$^{ \; a)}$\\
\vs{1cm}
{\it Universit\`a di Lecce, Dipartimento di Fisica\\
Instituto di Fisica Nucleare, Sezione di Lecce\\
Lecce, 73100, ITALY}
\vs{2cm}
{\bf Abstract}
\end{center}
We reformulate Special Relativity by a quaternionic algebra on reals.
Using {\em real linear quaternions}, we show that previous difficulties,
concerning the appropriate transformations on the $3+1$ space-time, may be
overcome. This implies that a complexified quaternionic version of
Special Relativity is a choice and not a necessity.
\vs{2cm}
\noindent{\footnotesize a) e-mail:} {\footnotesize \sl <EMAIL>}
\end{titlepage}
\pagebreak
\section{Introduction}
``{\sl The most remarkable formula in mathematics is:
\begin{equation}
e^{i\theta} = \cos{\theta} + i \sin{\theta} \; \; .
\end{equation}
This is our jewel. We may relate the geometry to the algebra by
representing complex numbers in a plane
\[ x+iy = re^{i\theta} \; \; . \]
This is the unification of algebra and geometry.}'' -
Feynman~\cite{fey}.\\
We know that a rotation of $\alpha$-angle around the $z$-axis, can be
represented by $e^{i\alpha}$, in fact
\[ e^{i\alpha}(x+iy) = re^{i(\theta + \alpha)} \; \; . \]
In 1843, Hamilton in the attempt to generalize the complex
field in order to describe the rotation
in the three-dimensional space, discovered quaternions\f{Quaternions, as
used in this paper, will always mean ``real quaternions''
\[q=a+ib+jc+kd \; \; \; \; , \; \; \; \;
a, \; b, \; c, \; d \; \in \; {\cal R} \; \; .\]}.\\
Today a rotation about an axis passing trough the origin and parallel to a
given unitary vector $\vec{u} \equiv (u_{x}, \; u_{y}, \; u_{z})$ by an
angle $\alpha$, can be obtained taking the transformation
\begin{equation}
e^{(iu_{x}+ju_{y}+ku_{z})\frac{\alpha}{2}} (ix+jy+kz)
e^{-(iu_{x}+ju_{y}+ku_{z})\frac{\alpha}{2}} \; \; .
\end{equation}
Therefore if we wish to represent rotations in the three-dimensional space
and complete ``{\sl the unification of algebra and geometry}'' we need
quaternions.
The quaternionic algebra has been expounded in a series of papers~\cite{pap}
and books~\cite{boo} with particular reference to quantum mechanics, the reader
may refer to these for further details. For convenience we repeat and
develop the relevant points in the following section, where the terminology
is also defined.
Nothing that $U(1, \; q)$ is algebraically isomorphic to $SU(2, \; c)$, the
imaginary units $i, \; j, \; k$ can be realized by means of the
$2\times 2$ Pauli matrices through
\[ (i, \; j, \; k) \leftrightarrow (i\sigma_{3}, \; -i\sigma_{2}, \;
-i\sigma_{1}) \; \; . \]
\begin{center}
{\footnotesize (this particular representation of the imaginary units $i, \; j, \; k$
has been introduced in ref.~\cite{del}).}
\end{center}
So a quaternion $q$ can be represented by a $2\times 2$ complex matrix
\bel{a}
q \; \; \leftrightarrow \; \; {\cal Q} = \left(
\begin{array}{cc} z_{1} & -z_{2}^{*}\\
z_{2} & z_{1}^{*}\end{array} \right) \; \; ,
\end{equation}
where
\[ z_{1}=a+ib \; \; \; , \; \; \; z_{2}=c-id \; \; \; \; \in \; \;
{\cal C}(1, \; i) \; \; .\]
It follows that a quaternion with unitary norm is identified by a
unitary $2\times 2$ matrix with unit determinant, this gives the
correspondence between unitary quaternions\f{In a recent
paper~\cite{del2} the representation theory of the group $U(1, \; q)$ has been
discussed in detail.} $U(1, \; q)$ and $SU(2, \; c)$.\\
Let us consider the transformation law of a spinor (two-dimensional
representations of the rotation group)
\begin{equation}
{\psi}'={\cal U} \psi \; \; ,
\end{equation}
where
\[ \psi=\left( \begin{array}{c} z_{1}\\ z_{2}\end{array} \right) \; \;
\; \; , \; \; \; \; {\cal U} \in SU(2, \; c) \; \; . \]
We can immediately verify that
\[ \tilde{\psi}=\left( \begin{array}{c} -z_{2}\\ z_{1}\end{array} \right) \]
transforms as follows
\begin{equation}
{\tilde{\psi}}'={\cal U}^{*}\tilde{\psi} \; \; ,
\end{equation}
so
\[\left( \begin{array}{cc} z_{1} & -z_{2}^{*}\\
z_{2} & z_{1}^{*}\end{array} \right)' = {\cal U}
\left( \begin{array}{cc} z_{1} & -z_{2}^{*}\\
z_{2} & z_{1}^{*}\end{array} \right) \]
represents again the transformation law of a spinor.\\
Thanks to the identification~(\ref{a}) we can write the previous
transformations by real quaternions as follows
\[ q'={\cal U}q \; \; , \]
with $q=z_{1}+jz{_2}$ and $\cal U$ quaternion with unitary norm
{\footnotesize ($N({\cal U})={\cal U}^{+}{\cal U}=1$)}. Note
that we don't need right operators to indicate the transformation law of a
spinor.
Now we can obtain the transformation law of a three-dimensional
vector $\vec{r}\equiv (x, \; y, \; z)$ by product of spinors, in fact
if we consider the purely imaginary quaternion
\[ \omega=qiq^{+}=ix+jy+kz \]
or the corresponding traceless $2\times 2$ complex matrix
\[ \Omega =\psi i \psi^{+} = \left( \begin{array}{cc} ix & -y-iz\\
y-iz & -ix\end{array} \right) \; \; , \]
a rotation in the three-dimensional space can be written as
follows\f{``{\sl No such `trick' works to relate the full
four-vector $(ct, \; x, \; y, \; z)$ with real
quaternions.}'' - Penrose~\cite{pen}.}
\begin{center}
\begin{tabular}{cl}
${\omega}'={\cal U}\omega \; {\cal U}^{+}$ &
{\footnotesize (quaternions)} ,\\ \\
${\Omega}'={\cal U}\Omega \; {\cal U}^{+}$ & {\footnotesize ($2\times 2$ complex matrices)} .
\end{tabular}
\end{center}
For infinitesimal transformations, ${\cal U}=1+\vec{Q}\cdot \vec{\theta}$,
we find
\[ \vec{Q}\cdot \vec{r} \; ' = \vec{Q}\cdot \vec{r}+
\vec{Q}\cdot \vec{\theta} \; \vec{Q}\cdot \vec{r}-
\vec{Q}\cdot \vec{r} \; \vec{Q}\cdot \vec{\theta} \; \; ,\]
where
\[ \vec{Q}\equiv (i, \; j, \;k) \; \; \; , \; \; \;
\vec{\theta}\equiv (\alpha, \; \beta, \; \gamma) \; \; .\]
If we rewrite the mentioned above transformation in the following
form\f{{\em Barred} operators ${\cal O}\mid q$ act on quaternionic objects
$\Phi$ as in
\[({\cal O}\mid q)\Phi={\cal O} \Phi q \; \;. \]}
\begin{equation}
\vec{Q}\cdot \vec{r} \; ' = [1+\vec{\theta}\cdot (\vec{Q} -1 \mid \vec{Q})]
\; \vec{Q}\cdot \vec{r} \; \; ,
\end{equation}
we identify
\[ \frac{i - 1\mid i}{2} \; \; \; , \; \; \;
\frac{j - 1\mid j}{2} \; \; \; , \; \; \;
\frac{k - 1\mid k}{2} \; \; \; ,\]
as the generators for rotations in the three-dimensional space\f{The factor
$\frac{1}{2}$ guarantees that our generators satisfy the usual algebra:
\[ [A_{m}, \; A_{n}]=\epsilon_{mnp} A_{p} \; \; \; \; \; m, \; n, \; p =1, \; 2,
\; 3 \; \; .\]}.
Up till now, we have considered only particular operations on
quaternions. A quaternion $q$ can also be multiplied by unitary
quaternions $\cal V$ from the right. A possible transformation which
preserves the norm is given by
\bel{o4}
q'={\cal U} q \; {\cal V} \; \; ,
\end{equation}
\begin{center}
{\footnotesize $( \; {\cal U}^{+}{\cal U}={\cal V}^{+}{\cal V}=1 \; ) \; \; .$}
\end{center}
Since left and right multiplications commute, the group is locally isomorphic
to $SU(2)\times SU(2)$, and so to $O(4)$, the four-dimensional Euclidean
rotation group.
As far as here we can recognize only particular real linear quaternions,
namely
\[ 1 \; \; , \; \; i \; \; , \; \; j \; \; , \; \; k \; \; , \; \;
1 \; \vert \; i \; \; , \; \; 1 \; \vert \; j \; \; , \; \;
1 \; \vert \; k \; \; . \]
We haven't hope to describe the Lorentz group if we use only previous
objects. Analyzing the most general transformation on
quaternions (see section 4), we introduce new real linear quaternions which
allow us to overcome the above difficulty and so
obtain a quaternionic version of the Lorentz group, without the use of
complexified quaternions. This result
appears, to the best of our knowledge, for the first time in print.\\
First of that we briefly recall the standard way to rewrite special
relativity by a quaternionic algebra on complex (see section 3).\\
In section 5, we present a quaternionic version of the special group
$SL(2, \; c)$, which is as well-known collected to the Lorentz group.
Our conclusions are drawn in the final section.
\section{Quaternionic algebras}
A quaternionic algebra over a field $\cal F$ is a set
\[ {\cal H} = \{ \alpha + i \beta + j \gamma + k \delta \; \; \vert
\; \; \alpha , \; \beta , \; \gamma , \; \delta \; \in \; {\cal F} \} \]
with multiplication operations defined by following rules for imaginary
units $i, \; j, \; k$
\begin{eqnarray*}
i^{2}=j^{2}=k^{2} & = & -1 \; \; ,\\
jk=-kj & = & i \; \; , \\
ki=-ik & = & j \; \; , \\
ij=-ji & = & k \; \; .
\end{eqnarray*}
In our paper we will work with quaternionic algebras defined on reals and
complex, so in this section we give a panoramic review of such algebras.
We start with a quaternionic algebra on reals
\[ {\cal H}_{\cal R} = \{ \alpha + i \beta + j \gamma + k \delta \; \; \vert
\; \; \alpha , \; \beta , \; \gamma , \; \delta \; \in \; {\cal R} \} \; \; .\]
We introduce the quaternion conjugation denoted by $^{+}$ and defined by
\[ q^{+}=\alpha - i\beta -j\gamma - k\delta \; \; . \]
The previous definition implies
\[ (\psi \varphi)^{+} = \varphi^{+} \psi^{+} \; \; ,\]
for $\psi$, $\varphi$ quaternionic functions. A conjugation
operation which does not reverse the order of $\psi$, $\varphi$ factors is
given, for example, by
\[ \tilde{q} = \alpha - i\beta +j\gamma - k\delta \; \; . \]
An important difference between quaternions and complexified
quaternions is based on the concept of {\em division algebra}, which is a
finite dimensional algebra for which $a\neq 0$, $b\neq 0$ implies
$ab\neq 0$, in others words, which has not nonzero divisors of zero. A
classical theorem~\cite{bot} states that the only division algebras over
the reals are algebras of dimension $1$, $2$, $4$ and $8$; the only
associative algebras over the reals are $\cal R$, $\cal C$ and
${\cal H}_{\cal R}$ (Frobenius~\cite{fro}); the nonassociative division
algebras include the octonions $\cal O$ (but there are others as well; see
Okubo~\cite{oku}).\\
A simple example of a {\em nondivision} algebra is provided by the algebra
of complexified quaternions
\[ {\cal H}_{\cal C} = \{ \alpha + i \beta + j \gamma + k \delta \; \; \vert
\; \; \alpha , \; \beta , \; \gamma , \; \delta \; \in \;
{\cal C}(1, \; {\cal I}) \} \; \; ,\]
\[ [{\cal I}, \; i]=[{\cal I}, \; j]=[{\cal I}, \; k]=0 \; \; .\]
In fact since
\[ (1+i{\cal I})(1-i{\cal I})=0 \; \; ,\]
there are nonzero divisors of zero.
For complexified quaternions we have different opportunities to define
conjugation operations, we shall use the following terminology:
\begin{enumerate}
\item The {\sl complex} conjugate of $q_{\cal C}$ is
\[ q_{\cal C}^{\; *} = {\alpha}^{*} + i{\beta}^{*} +j{\gamma}^{*} +k{\delta}^{*} \; \; .
\]
Under this operation
\begin{eqnarray*}
({\cal I}, \; i, \; j, \; k) & \rightarrow & (-{\cal I}, \; i, \; j, \; k)
\; \; ,
\end{eqnarray*}
and
\[ (q_{\cal C}p_{\cal C})^{*}=q_{\cal C}^{\; *} p_{\cal C}^{\; *} \; \; .\]
\item The {\sl quaternion} conjugate of $q_{\cal C}$ is
\[ q_{\cal C}^{\; \star} = \alpha - i\beta -j\gamma -k \delta \; \; . \]
Here
\begin{eqnarray*}
({\cal I}, \; i, \; j, \; k) & \rightarrow & ({\cal I}, \; -i, \; -j, \; -k)
\; \; ,
\end{eqnarray*}
and
\[ (q_{\cal C}p_{\cal C})^{\star}=p_{\cal C}^{\; \star}
q_{\cal C}^{\; \star} \; \; .\]
\item In the absence of standard terminology, we call that formed by
combining these operations, the {\sl full} conjugate
\[ q_{\cal C}^{\; +} = {\alpha}^{*} - i{\beta}^{*} -j{\gamma}^{*} -k{\delta}^{*} \; \; .
\]
Under this operation
\begin{eqnarray*}
({\cal I}, \; i, \; j, \; k) & \rightarrow & -({\cal I}, \; i, \; j, \; k)
\; \; ,
\end{eqnarray*}
and
\[ (q_{\cal C}p_{\cal C})^{+}=p_{\cal C}^{\; +}
q_{\cal C}^{\; +} \; \; .\]
\end{enumerate}
\section{Complexified Quaternions and Special Relati\-vi\-ty}
We begin this section by recalling a sentence of Anderson
and Joshi~\cite{and} about the quaternionic reformulation of special
relativity:
``{\sl There has been a long tradition of using quaternions for Special
Relativity ... The use of quaternions in special relativity, however, is not
entirely straightforward. Since the field of quaternions is a
four-dimensional Euclidean space, complex components for the quaternions
are required for the $3+1$ space-time of special relativity}.''
In the following section, we will demonstrate
that a reformulation of special relativity by a quaternionic algebra on
reals is possible.
In the present section, we use complexified
quaternions to reformulate special relativity, for further details the reader
may consult the papers of Edmonds~\cite{edm}, Gough~\cite{gou},
Abonyi~\cite{abo}, G\"ursey~\cite{gur} and the book of Synge~\cite{syn}.
A space-time point can be represented by complexified quaternions as follows
\begin{equation}
{\cal X} = {\cal I}ct + ix+jy+kz \; \; .
\end{equation}
The Lorentz invariant in this formalism is given by
\begin{equation}
{\cal X}^{*}{\cal X}=(ct)^{2} - x^{2} - y^{2} - z^{2} \; \; .
\end{equation}
If we consider the standard Lorentz transformation (boost $ct$ - $x$)
\begin{eqnarray*}
ct' & = & \gamma \; (ct-\beta x) \; \; ,\\
x' & = & \gamma \; (x-\beta ct) \; \; ,\\
y' & = & y \; \; ,\\
z' & = & z \; \;
\end{eqnarray*}
and note that the first two equations may be rewritten as
\begin{eqnarray*}
ct' & = & ct \cosh{\theta} - x \sinh{\theta} \; \; ,\\
x' & = & x \cosh{\theta} - ct \sinh{\theta} \; \; ,
\end{eqnarray*}
\begin{center}
{\footnotesize where $\cosh{\theta} =\gamma \; \; \; , \; \; \;
\sinh{\theta} = \beta \gamma \; \; ,$}
\end{center}
we can represent an infinitesimal transformation by
\begin{eqnarray*}
{\cal X}' & = & {\cal I}(ct-x\theta ) +i(x -ct\theta )+jy+kz \; =\\
& = & {\cal X} + {\cal I} \; \frac{i+1\mid i}{2} \theta {\cal X} \; \; .
\end{eqnarray*}
We thus recognize, in the previous transformation, the generator
\[ {\cal I} \; \frac{i+1\mid i}{2} \; \; . \]
It is now very simple to complete the translation, the set of generators
of the Lorentz group is provided with
\begin{eqnarray*}
boost \; \; (ct, \; x) \; \; &
\; \; {\cal I} \; \frac{i+1 \; \vert \; i}{2} \; \; ,\\
boost \; \; (ct, \; y) \; \; &
\; \; {\cal I} \; \frac{j+1 \; \vert \; j}{2} \; \; ,\\
boost \; \; (ct, \; z) \; \; &
\; \; {\cal I} \; \frac{k+1 \; \vert \; k}{2} \; \; ,\\
rotation \; \; around \; \; x \; \; &
\; \; \; \; \; \frac{i-1 \; \vert \; i}{2} \; \; ,\\
rotation \; \; around \; \; y \; \; &
\; \; \; \; \; \frac{j-1 \; \vert \; j}{2} \; \; ,\\
rotation \; \; around \; \; z \; \; &
\; \; \; \; \; \frac{k-1 \; \vert \; k}{2} \; \; .
\end{eqnarray*}
And so a general finite Lorentz transformation is given by
\[ e^{{\cal I}(i\alpha_{b}+j\beta_{b}+k\gamma_{b})+
i\alpha_{r} +j\beta_{r} +k \gamma_{r}} ({\cal I}ct + ix+jy+kz)
e^{{\cal I}(i\alpha_{b}+j\beta_{b}+k\gamma_{b})-
i\alpha_{r} -j\beta_{r} -k \gamma_{r}} \; \; . \]
The previous results can be elegantly summarize by the relation
\begin{equation}
{\cal X}'=\Lambda {\cal X} \Lambda^{+} \; \; \; \; , \; \; \; \;
\Lambda^{\star} \Lambda = 1 \; \; ,
\end{equation}
where $\Lambda$ is obviously a complexified quaternion. In this or similar way
a lot of authors have reformulated special relativity with complex
quaternions.
In the following section we will introduce {\em real linear quaternions}
and reformulate special relativity using a quaternionic algebra on
reals. We remark that complex component for the quaternions represent a
choice and not a necessity.
\section{A new possibility}
We think that quaternions are the natural candidates to
describe special relativity. It is simple to understand why,
quaternions are characterized by four real numbers (whereas complexified
quaternions by eight), thus we can collect these four
real quantities with a point $(ct, \; x, \; y, \; z)$ in the space-time.
In quaternionic notation we have
\begin{equation}
{\cal X}= ct + ix + jy + kz \; \; .
\end{equation}
In the first section we have introduced particular {\em real linear
quaternions}, namely
\[ 1 \; \; , \; \; \vec{Q} \; \; , \; \; 1\mid \vec{Q} \; \; ,\]
where
\[ \vec{Q}\equiv (i, \; j, \; k) \; \; . \]
In order to write the most general real linear quaternions we must consider
the following quantities
\[ \vec{Q}\mid i \; \; , \; \; \vec{Q}\mid j \; \; , \; \; \vec{Q}\mid k
\; \; .\]
In fact the most general transformation on quaternions is represented by
\bel{w}
q+ p\mid i + r\mid j + s\mid k \; \; ,
\end{equation}
with
\[ q, \; p, \; r, \; s \;
\in \; {\cal H}_{\cal R} \; \; . \]
New objects like
\[ k\mid j \; \; , \; \; j\mid k \; \; , \; \; i\mid k \; \; , \; \;
k\mid i \; \; , \; \; j\mid i \; \; , \; \; i\mid j \]
will be essential to reformulate special relativity with real quaternions.
They represent the wedge which permit to overcome the difficulties which
in past did not allow a (real) quaternionic version of special relativity.
Returning to Lorentz transformations, let us start with the following
infinitesimal transformation (boost $ct$ - $x$)
\begin{eqnarray*}
{\cal X}' & = & ct-x\theta +i(x -ct\theta )+jy+kz \; =\\
& = & {\cal X} + \frac{k\mid j - j\mid k}{2} \; \theta {\cal X} \; \; .
\end{eqnarray*}
We can immediately note that the generator which substitutes
\[ {\cal I} \; \frac{i+1\mid i}{2} \]
is
\[ \frac{k\mid j - j\mid k}{2} \; \; . \]
So we have (for the first time in print) the possibility to list the
generators of the Lorentz group without the need to work with complexified
quaternions
\begin{eqnarray*}
boost \; \; (ct, \; x) \; \; &
\; \; \frac{k \; \vert \; j - j \; \vert \; k}{2} \; \; ,\\
boost \; \; (ct, \; y) \; \; &
\; \; \frac{i \; \vert \; k - k \; \vert \; i}{2} \; \; ,\\
boost \; \; (ct, \; z) \; \; &
\; \; \frac{j \; \vert \; i - i \; \vert \; j}{2} \; \; ,\\
rotation \; \; around \; \; x \; \; &
\; \; \; \; \frac{i-1 \; \vert \; i}{2} \; \; ,\\
rotation \; \; around \; \; y \; \; &
\; \; \; \; \frac{j-1 \; \vert \; j}{2} \; \; ,\\
rotation \; \; around \; \; z \; \; &
\; \; \; \; \frac{k-1 \; \vert \; k}{2} \; \; .
\end{eqnarray*}
In appendix A we explicitly prove that the action of previous generators
leaves
\begin{equation}
Re \; {\cal X}^{2} = (ct)^{2} - x^{2} - y^{2} - z^{2}
\end{equation}
invariant.
In appendix B we will give an alternative but equivalent presentation of
special relativity by a quaternionic algebra on reals. There we introduce a
real linear quaternion $g$ which substitutes the metric tensor $g^{\mu \nu}$.
\section{A quaternionic version of the complex group $SL(2)$}
In analogy to the connection between the rotation group
$O(3)$ to the special unitary group $SU(2)$, there is a natural
correspondence~\cite{tun} between
the Lorentz group $O(3,~1)$ and the special linear group
$SL(2)$. In fact $SL(2)$ is the universal covering group of $O(3,~1)$ in
the same way that $SU(2)$ is of $O(3)$.
The aim of this Section is to give, by extending the consideration which
collect the special unitary group $SU(2)$ with unitary real quaternions
(as shown in section 1), a quaternionic version of
the special linear group $SL(2)$.
Once more the aim will be achieved with help of real linear quaternions.
A Lorentz spinor is a complex object which transform under Lorentz
transformations as
\[ \psi' = {\cal A} \psi \; \; , \]
where $\cal A$ is a $SL(2)$ matrix. When we restrict ourselves to the
three-dimensional space and to rotations, this definition gives the usual
Pauli spinors
\[ \psi' = {\cal U} \psi \; \; , \]
where $\cal U$ is a $SU(2)$ matrix.
Now we shall derive the generators of rotations and Lorentz boosts in the
spinor representation by using real linear quaternions.
The action of generators of the special group $SL(2)$
\begin{eqnarray*}
\left( \begin{array}{cc} $i$ & 0\\ 0 & $-i$\end{array} \right) \; , &
\left( \begin{array}{cc} 0 & $-1$\\ 1 & 0\end{array} \right) \; , &
\left( \begin{array}{cc} 0 & $-i$\\ $-i$ & 0\end{array} \right) \; ,\\ \\
\left( \begin{array}{cc} $-1$ & 0\\ 0 & 1\end{array} \right) \; , &
\left( \begin{array}{cc} 0 & $-i$\\ $i$ & 0\end{array} \right) \; , &
\left( \begin{array}{cc} 0 & 1\\ 1 & 0\end{array} \right) \; ,
\end{eqnarray*}
on the spinor
\[ \psi = \left( \begin{array}{c} \xi \\ \eta \end{array} \right) \; , \]
can be represented by the action of real linear quaternions
\begin{eqnarray*}
i \; , & j \; , & k \; ,\\
i \; \vert \; i \; , & j \; \vert \; i \; , & k \; \vert \; i \; ,
\end{eqnarray*}
on the quaternion
\[ q= \xi +j\eta \; \; . \]
In section 1 we have obtained a three-dimensional vector $(x, \; y, \; z)$
by product of Pauli spinors $q_{\cal P}$:
\[q_{\cal P}\; i \; q_{\cal P}^{+}=ix+jy+kz \]
\begin{center}
{\footnotesize $ ( \; q'_{\cal P}={\cal U}q_{\cal P} \; \; \; , \; \; \;
{\cal U}^{+}{\cal U}=1 \; ) \; \; ,$}
\end{center}
consequently we have written its transformation law as follows
\[(q_{\cal P} \; i \; q_{\cal P}^{+})'={\cal U}q_{\cal P} \; i \;
q_{\cal P}^{+} \; {\cal U}^{+}
\; \; .\]
Now we start with a Lorentz spinor $q_{\cal L}$
\[q'_{\cal L}={\cal A}q_{\cal L} \; \; ,\]
and construct a four-vector $(ct, \; x, \; y, \; z)$ by product of such
spinors
\[q_{\cal L} \; (1+i) \; q_{\cal L}^{+}=ct+ix+jy+kz \; \; .\]
The transformation law is then given by
\[(q_{\cal L} \; (1+i) \; q_{\cal L}^{+})'=({\cal A}q_{\cal L}) \; (1+i)
\; ({\cal A}q_{\cal L})^{+} \; \; .\]
If we consider infinitesimal transformations
\[ {\cal A} = 1+\frac{\vec{Q}}{2}\cdot (\vec{\theta}+\vec{\zeta}\mid i) \; \; ,\]
\begin{center}
{\footnotesize with $\vec{\theta}\equiv (\alpha, \; \beta, \; \gamma)$ and
$\vec{\zeta}\equiv (\tilde{\alpha}, \; \tilde{\beta}, \; \tilde{\gamma}) \; \; ,$}
\end{center}
we have
\begin{eqnarray*}
{\cal T}' & = & {\cal T}+\frac{\alpha}{2}[i, \; {\cal T}]+
\frac{\beta}{2}[j, \; {\cal T}]+\frac{\gamma}{2}[k, \; {\cal T}]+\\
& & + \; \frac{\tilde{\alpha}}{2} \{i, \; \tilde{{\cal T}}\}+
\frac{\tilde{\beta}}{2} \{j, \; \tilde{{\cal T}}\}+
\frac{\tilde{\gamma}}{2}\{k, \; \tilde{{\cal T}} \}
\end{eqnarray*}
where
\[ {\cal T}= q_{\cal L} \; (1+i) \; q_{\cal L}^{+} \; \; ,\]
and
\[ \tilde{{\cal T}} =q_{\cal L} \; i(1+i) \; q_{\cal L}^{+}=
{\cal T} -2q_{\cal L}q_{\cal L}^{+}\; \; .\]
In order to simplify next considerations we pose
\begin{eqnarray*}
{\cal T}=ix+jy+kz+ct & = {\cal T}_{i}+{\cal T}_{j}+{\cal T}_{k}+
{\cal T}_{1} \; \; , \\
\tilde{\cal T}=ix+jy+kz-ct & = {\cal T}_{i}+{\cal T}_{j}+{\cal T}_{k}-
{\cal T}_{1} \; \; ,
\end{eqnarray*}
so the standard Lorentz transformations are given by
\begin{eqnarray*}
{\cal T}_{1} & \rightarrow & {\cal T}_{1} +\tilde{\alpha} i {\cal T}_{i}+
\tilde{\beta} j {\cal T}_{j}+\tilde{\gamma} k {\cal T}_{k}\\
{\cal T}_{i} & \rightarrow & {\cal T}_{i} -\tilde{\alpha} i {\cal T}_{1}+
\beta j {\cal T}_{k}-\gamma k {\cal T}_{j}\\
{\cal T}_{j} & \rightarrow & {\cal T}_{j} -\tilde{\beta} j {\cal T}_{1}-
\alpha i {\cal T}_{k}+\gamma k {\cal T}_{i}\\
{\cal T}_{k} & \rightarrow & {\cal T}_{k} -\tilde{\gamma} k {\cal T}_{1}+
\alpha i {\cal T}_{j}-\beta j {\cal T}_{i} \; \; .
\end{eqnarray*}
In this way we obtain a quaternionic version of the special
group $SL(2)$ and demonstrate\f{In contrast with the opinion of
Penrose~\cite{pen}, cited in footnote 3.} that, if real linear
quaternions appear, a `trick' similar to that one of rotations works to
relate the full four-vector $(ct, \; x, \; y, \; z)$ with real
quaternions.
\section{Conclusions}
The study of special relativity with a quaternionic algebra on reals has
yielded a result of interest. While we cannot demonstrate in this paper
that one number system (quaternions) is preferable to another
(complexified quaternions) we have pointed out the advantages of using
real linear quaternions which naturally appear when we work with a non
commutative number system, like the quaternionic field. As seen in this paper
these objects are very useful if we wish to rewrite special relativity by a
quaternionic algebra on reals. The complexified
quaternionic reformulation of special relativity is thus a choice and
not a necessity. This affirmation is in contrast with the standard folklore
(see, for example, Anderson and Joshi~\cite{and}).
Our principal aim in this work is to underline the potentialities of real
linear quaternions. We wish to remember that a lot of difficulties have
been overcome thanks to these objects (which in our colourful
language we have named generalized objects~\cite{del}).\\
To remark their potentialities let us list the situations which have
requested their use.
\begin{itemize}
\item The need of such objects naturally appears, for
example, in the construction of quaternion group theory and tensor
product group representations~\cite{del2}. Also starting with only standard
quaternions $i, \; j, \; k$ in order to represent the generators of
the group $U(1, \; q)$, we find generalized quaternions when we analyze
quaternionic tensor products.
\[ Spin \; \; \frac{1}{2} \; \; generators \; : \; \;
\frac{i}{2} \; \; , \; \; \frac{j}{2} \; \; , \; \; \frac{k}{2} \; \; .\]
\[ Spin \; \; 1\oplus 0 \; \; generators \; :\]
\[ \left( \begin{array}{cc} \frac{i+1 \; \vert \; i}{2} & 0\\
0 & \frac{i-1 \; \vert \; i}{2} \end{array} \right) \; , \;
\left( \begin{array}{cc} j & 1 \; \vert \; i\\
1 \; \vert \; i & j\end{array} \right) \; , \;
\left( \begin{array}{cc} k & -1\\
1 & k\end{array} \right) \; \; .\]
\item If we desire to extend the isomorphism of $SU(2, \; c)$ with
$U(1, \; q)$ to the group $U(2, \; c)$, we must introduce the additional
real linear quaternion `$1 \; \vert \; i$'.
In this way there exists at least one version
of quaternionic quantum mechanics in which a `partial' set of translations
may be defined~\cite{del}, in fact, tanks to real linear operators, a
translation between $2n\times 2n$ complex and $n\times n$ quaternionic
matrices is possible.
\item In the work of ref.~\cite{rot} a quaternion version of the Dirac
equation was derived in the form
\[ \gamma_{\mu} \partial^{\mu} \psi i = m \psi \; \; ,\]
where the $\gamma_{\mu}$ are two by two quaternionic matrices satisfying the
Dirac condition
\[ \{\gamma_{\mu}, \; \gamma_{\nu} \} = 2 g_{\mu \nu} \; \; . \]
In the Rotelli's formalism the momentum operator must be defined as
\[ p^{\mu} = \partial^{\mu} \; \vert \; i \]
which is also a generalized object.
\item In this paper, contrary to the common opinion, we have given a real
quaternionic formulation of special relativity. In order to obtain that we have
introduced the following real linear quaternions
\[ \vec{Q} \; \vert \; i \; , \; \vec{Q} \; \vert \; j \; ,
\; \vec{Q} \; \vert \; k \; \; , \]
\[ \vec{Q} \equiv (i, \; j, \; k) \; \; . \]
A quaternionic version of the special group $SL(2)$ has also been given.
\end{itemize}
We finally note that the process of generalization can be extend also
to complexified quaternions. In a recent paper~\cite{del3} we give an
elegant one-component formulation of the Dirac equation and, thanks to our
generalization, we overcome previous
difficulties concerning the doubling of solutions~\cite{and,edm,gou} in the
complexified quaternionic Dirac equation.
In seeking a better understanding of the success of mathematical
abstraction in physics and in particular of the wide applicability of
quaternionic numbers in theories of physical phenomena, we found that
generalized quaternions should be {\em not} undervalued. We think
that there are good reasons to hope that these generalized structures provide
new possibilities concerning physical applications of quaternions.
``{\sl The most powerful method of advance that can be suggested at present
is to employ all the resources of pure mathematics in attempts to perfect
and generalize the mathematical formalism that forms the existing basis of
theoretical physics, and after each success in this direction, to try to
interpret the new mathematical features in terms of physical entities...}''
- Dirac~\cite{dir}.
\section*{Appendix A}
In this appendix we prove that the Lorentz invariant is
\bel{apb}
Re \; {\cal X}'^{\; 2}=Re \; {\cal X}^{2} \; \; ,
\end{equation}
where
\[{\cal X}=ct+ix+jy+kz \; \; .\]
Under an infinitesimal transformation, we have
\[ {\cal X}' = (1+\theta \; \frac{k \; \vert \; j-j \; \vert \; k}{2}+
\alpha \; \frac{i-1 \; \vert \; i}{2}+...){\cal X} \]
so, neglecting second order terms,
\[ {\cal X}'^{\; 2} = {\cal X}^{2}+\frac{\theta}{2} \; \{{\cal X}, \;
k{\cal X}j-j{\cal X}k \}+
\frac{\alpha}{2} \; \{{\cal X}, \; i{\cal X}-{\cal X}i \}+ ... \]
Equation~(\ref{apb}) is then satisfied since
\begin{eqnarray*}
\{{\cal X}, \; i{\cal X}-{\cal X}i \} & = & (i-1\mid i) {\cal X}^{2} \; \;
,\\
\{{\cal X}, \; k{\cal X}j-j{\cal X}k \} & = & (1\mid j -j) {\cal X}k{\cal X} +
(k - 1\mid k) {\cal X}j{\cal X} \; \; ,
\end{eqnarray*}
are purely imaginary quaternions.
Obviously we can derive the generators of Lorentz group by starting
from the infinitesimal transformation
\[ {\cal X}'={\cal X}+{\cal A}{\cal X} \]
and imposing that they satisfy the
relation
\begin{equation}
Re \; \{{\cal X}, \; {\cal A}{\cal X} \}=0
\end{equation}
\begin{center}
{\footnotesize $( \; Re \; {\cal X}'^{\; 2}=Re \; {\cal X}^{2} \; \;
\Rightarrow \; \; Re \; \{{\cal X}, \; {\cal A}{\cal X} \}=0 \; )$ .}
\end{center}
With straightforward mathematical calculus we can find the generators
requested. In order to simplify following considerations let us pose
\[ {\cal X}=a+ib+jc+kd \; \; \; , \; \; \;
{\cal A}=q_{0}+q_{1}\mid i+q_{2}\mid j+q_{3}\mid k\]
\begin{center}
{\footnotesize where $q_{m}={\alpha}_{m}+i{\beta}_{m}+j{\gamma}_{m}+k{\delta}_{m} \; \;
(m \; = \; 0, \; 1, \; 2, \; 3)$ are real quaternions.}
\end{center}
The only quantities which we must calculate are
\[ Re \; \{{\cal X}, \; {\cal X} \} \; \; , \; \;
Re \; \{{\cal X}, \; i{\cal X}i \} \; \; , \; \;
Re \; \{{\cal X}, \; i{\cal X} \} \; \; , \; \;
Re \; \{{\cal X}, \; k{\cal X}j \} \; \; ,\]
in fact the other quantities can be obtained from previous ones, by simple
manipulations.
\begin{eqnarray*}
Re \; \{{\cal X}, \; {\cal X} \} \; \; =2(+a^{2}-b^{2}-c^{2}-d^{2}) & , &
Re \; \{{\cal X}, \; i{\cal X}i \} \; =2(-a^{2}+b^{2}-c^{2}-d^{2})\\
Re \; \{{\cal X}, \; j{\cal X}j \}=2(-a^{2}-b^{2}+c^{2}-d^{2}) & , &
Re \; \{{\cal X}, \; k{\cal X}k \}=2(-a^{2}-b^{2}-c^{2}+d^{2})\\
Re \; \{{\cal X}, \; i{\cal X} \}=Re \; \{{\cal X}, \; {\cal X}i \}=-4ab & , &
Re \; \{{\cal X}, \; k{\cal X}j \}=Re \; \{{\cal X}, \; j{\cal X}k \}=4cd\\
Re \; \{{\cal X}, \; j{\cal X} \}=Re \; \{{\cal X}, \; {\cal X}j \}=-4ac & , &
Re \; \{{\cal X}, \; j{\cal X}i \} \; =Re \; \{{\cal X}, \; i{\cal X}j \} \;
=4bc\\
Re \; \{{\cal X}, \; k{\cal X} \}=Re \; \{{\cal X}, \; {\cal X}k \}=-4ad & , &
Re \; \{{\cal X}, \; i{\cal X}k \}\; =Re \; \{{\cal X}, \; k{\cal X}i \} \;
=4bd \; \; .
\end{eqnarray*}
The previous relations imply the following conditions on the real parameters
of the generator $\cal A$
\begin{eqnarray*}
\alpha_{0} = 0 & , & \beta_{1} = 0 \\
\gamma_{2} = 0 & , & \delta_{3} = 0 \\
\beta_{0} = - \alpha_{1} = \alpha & , & \gamma_{0} = - \alpha_{2} = \beta \\
\delta_{0} = - \alpha_{3} = \gamma & , & \delta_{2} = - \gamma_{3} = \theta \\
\gamma_{1} = - \beta_{2} = \varphi & , & \beta_{3} = - \delta_{1} = \eta \; \;
{}.
\end{eqnarray*}
We can immediately recognize the Lorentz generators given in section 4.
\section*{Appendix B}
We introduce the usual four-vector $x^{\mu}$ by the following
quaternion
\[ {\cal X} = x^{0} + i x^{1} + j x^{2} + k x^{3} \; \; , \]
and define a scalar product of two vectors $\cal X$, $\cal Y$ by
\begin{equation} \label{rsp}
({\cal X}, \; g{\cal Y})_{\cal R} = Re \; ({\cal X}^{+}g{\cal Y}) =
x^{\mu} g_{\mu \nu} y^{\nu} \; \; ,
\end{equation}
where $g$ is the generalized quaternion
\[ - \frac{1}{2} \; (1 + i \; \vert \; i + j \; \vert \; j +
k \; \vert \; k) \; \; .\]
We can define a real norm (or metric)
\[ ({\cal X}, \; g{\cal X})_{\cal R} = Re \; ({\cal X}^{+}g{\cal X}) =
x^{\mu} g_{\mu \nu} x^{\nu} \; \; .\]
The vectors which transform under a Lorentz transformation ${\cal L}$, will be
denoted by
\[ {\cal X}' = {\cal L}{\cal X} \; \; , \]
with $\cal L$ real linear operators (see eq.~(\ref{w})).\\ From the
postulated invariance of the norm we can deduce the generators of Lorentz
group.
If we consider infinitesimal transformations
\[ {\cal L} = 1 + {\cal A} \; \; ,\]
we have
\[ Re \; ({\cal X}'^{+}g{\cal X}') = Re \; ({\cal X}^{+}g{\cal X} +
{\cal X}^{+}( {\cal A}^{+}g + g{\cal A}){\cal X}) =
Re \; ({\cal X}^{+}g{\cal X})
\; \; ,\]
and therefore
\bel{gen}
{\cal A}^{+}g + g{\cal A} = 0 \; \; .
\end{equation}
Using real scalar products, given an operator
\[ {\cal A}=q+p \; \vert \; i+r \; \vert \; j+s \; \vert \; k \; \; ,\]
\[ q, \; p, \; r, \; s \; \in \; {\cal H}_{\cal R} \; \; ,\]
we can write its hermitian conjugate as follows
\[ {\cal A}^{+}=q^{+}-p^{+} \; \vert \; i-r^{+} \; \vert \; j-
s^{+} \; \vert \; k \; \; .\]
Then the equation~(\ref{gen}) can be rewritten as
\[ g{\cal A}+h.c.=0 \; \; .\]
If we pose
\[ g{\cal A}=B= \tilde{q}+ \tilde{p} \; \vert \; i+\tilde{r} \; \vert \; j+
\tilde{s} \; \vert \; k \; \; ,\]
we obtain the following conditions on the operator $B$
\[ Re \; \tilde{q} = Vec \; \tilde{p} = Vec \; \tilde{r} =
Vec \; \tilde{s} = 0 \; \; .\]
Noting that ${\cal A}=gB$ we can quickly write the generators of
Lorentz group. We give explicitly an example
\[ {\cal A}_{1}=g \; (1 \; \vert \; i) = -\frac{1}{2}(-i+1 \; \vert \; i +
j \; \vert \; k - k \; \vert \; j) \; \; , \]
\[ {\cal A}_{2}=g \; i = -\frac{1}{2}(i-1 \; \vert \; i +j \; \vert \; k -
k \; \vert \; j) \; \; ,\]
\begin{eqnarray*}
{\cal A}={\cal A}_{1}-{\cal A}_{2} & = & \frac{i - 1 \; \vert \; i}{2} \; \;
,\\
\tilde{{\cal A}}={\cal A}_{1}+{\cal A}_{2} & = &
\frac{k \; \vert \; j - j \; \vert \; k}{2} \; \; .
\end{eqnarray*}
\begin{thebibliography}{99}
\bibitem{fey}
{\em Feynman RP} - {\sl The Feynman Lectures on Physics}, vol.~I - part 1.\\
Inter European Editions, Amsterdam (1975).
\bibitem{pap}
{\em Finkelstein D, Jauch JM, Schiminovich S, Speiser D} - \\
J.~Math.~Phys. {\bf 3}, 207 (1962); {\bf 4}, 788 (1963).\\
{\em Finkelstein D, Jauch JM, Speiser D} - J.~Math.~Phys. {\bf 4}, 136
(1963).\\
{\em Adler SL} - Phys.~Rev. {\bf D21}, 550 (1980), {\bf D21}, 2903 (1980);\\
{\em Adler SL} - Phys.~Rev.~Letts {\bf 55}, 783 (1985);\\
{\em Adler SL} - Phys.~Rev. {\bf D34}, 1871 (1986);\\
{\em Adler SL} - Comm.~Math.~Phys. {\bf 104}, 611 (1986);\\
{\em Adler SL} - Phys.~Rev. {\bf D37}, 3654 (1988);\\
{\em Adler SL} - Nuc.~Phys. {\bf B415}, 195 (1994).\\
\pr{De Leo S, Rotelli P}{D45}{575}{92};\\
{\em De Leo S, Rotelli P} - {\sl Quaternions Higgs and Electroweak Gauge
Group} - {\footnotesize to appear in {\bf Int.~J.~of Mod.~Phys.~A}}.\\
{\em De Leo S} - {\sl Quaternions for GUTs} - {\footnotesize submitted for
publication in {\bf J.~Math.~Phys.}}\\
{\em Dimitir\'c R, Goldsmith B} - Math.~Intell. {\bf 11}, 29 (1989).\\
{\em Horwitz LP, Biedenharn LC} - Ann.~of Phys. {\bf 157}, 432 (1984).\\
{\em Horwitz LP} -J.~Math.~Phys. {\bf 34}, 3405 (1993).\\
{\em Razon A, Horwitz LP} - Acta Appl.~Math. {\bf 24}, 141 (1991); 179
(1991);\\
{\em Razon A, Horwitz LP} - J.~Math.~Phys. {\bf 33}, 3098 (1992).\\
{\em Rembieli\'nski J} - J.~Phys. {\bf A11}, 2323 (1978),
{\bf A13}, 15 (1980);\\
{\em Rembieli\'nski J} - J.~Phys. {\bf A13}, 23 (1980),
{\bf A14}, 2609 (1981).\\
{\em Nash CC and Joshi GC} - J.~Math.~Phys. {\bf 28}, 2883, 2886 (1987);\\
{\em Nash CC and Joshi GC} - Int.~J.~Theor.~Phys. {\bf 27}, 409 (1988);\\
{\em Nash CC and Joshi GC} - Int.~J.~Theor.~Phys. {\bf 31}, 965 (1992).
\bibitem{boo}
{\em Finkelstein D, Jauch JM, Speiser D} - {\sl Notes on quaternion quantum
mechanics}\\
{\footnotesize in Logico-Algebraic Approach to Quantum Mechanics II}, Hooker (Reidel,
Dordrecht 1979), 367-421.\\
{\em Hamilton WR} - {\sl Elements of Quaternions} - Chelsea Publishing Co.,
N.Y., 1969.\\
{\em Gilmore R} - {\sl Lie Groups, Lie Algebras and Some of their
Applications} - John Wiley \& Sons, 1974.\\
{\em Altmann SL} - {\sl Rotations, Quaternions, and Double Groups} - Claredon,
Oxford, 1986.\\
{\em Adler SL} - {\sl Quaternion quantum mechanics and quantum fields} -
Oxford UP, Oxford, 1995.
\bibitem{del}
{\em De Leo S, Rotelli P} - Prog.~Theor.~Phys. {\bf 92}, 917 (1994).
\bibitem{del2}
\nc{De Leo S, Rotelli P}{B110}{33}{95}.
\bibitem{pen}
{\em Penrose R, Rindler W} - {\sl Spinors and space-time} - Cambridge UP,
Cambridge, 1984 ( vol.~1, pag.~23).
\bibitem{bot}
{\em Bott R, Milnor J} - Bull.~Amer.~Math.~Soc. {\bf 64}, 87 (1958).\\
{\em Kervaire M} - Proc.~Nat.~Acad.~Sci. {\bf 44}, 280 (1958).
\bibitem{fro}
{\em Frobenius G} - J.~Reine Angew.~Nath. {\bf 84}, 59 (1878).
\bibitem{oku}
{\em Okubo S} - {\sl Introduction to Octonion and Other Non-Associative
Algebras in Physics} - {\footnotesize unpublished}.
\bibitem{and}
{\em Anderson R, Joshi GC} - Phys.~Essays {\bf 6}, 308 (1993).
\bibitem{edm}
{\em Edmonds JD} - Int.~J.~Theor.~Phys. {\bf 6}, 205 (1972).\\
{\em Edmonds JD} - Found.~of Phys. {\bf 3}, 313 (1973).\\
{\em Edmonds JD} - Am.~J.~Phys. {\bf 42}, 220 (1974).
\bibitem{gou}
\ejp{Gough W}{7}{35}{86}; \xx{8}{164}{87}; \xx{10}{188}{89}.
\bibitem{abo}
{\em Abonyi I, Bito JF, Tar JK} - J.~Phys.~A {\bf 24}, 3245 (1991).
\bibitem{gur}
{\em G\"ursey F} - {\sl Symmetries in Physics (1600-1980): Proceedings of the
$1^{st}$ International Meeting on the History of Scientific Ideas} -
Seminari d'~Historia de les Ciences, Barcellona, Spain, 557 (1983).
\bibitem{syn}
{\em Synge JL} - {\sl Quaternions, Lorentz Transformation, and the Conway
Dirac Eddington Matrices} - Dublin Institute for Advanced Studies, Dublin,
1972.
\bibitem{tun}
{\em Wu-Ki Tung} - {\sl Group Theory in Physics} - World Scientific,
Singapore, 1985.
\bibitem{rot}
{\em Rotelli P} - Mod.~Phys.~Lett.~A {\bf 4}, 933 (1989).
\bibitem{del3}
{\em De Leo S} - {\sl One-component Dirac equation} - {\footnotesize submitted for
publication in {\bf Int.~J.~of Phys.~A}.}
\bibitem{dir}
{\em Dirac PAM} - Proc.~R.~Soc.~London {\bf A133}, 60 (1931).
\end{thebibliography}
\end{document}
|
\section{Introduction}
Heavy-ion collisions at energies in the vicinity of the nuclear
Coulomb barrier
lead to an alignment of the colliding nuclei. This implies
that the magnetic
substates are no longer equally populated. To describe deexcitation
processes following heavy-ion collisions such as $\gamma$-ray emission or
internal conversion, we have to account for this
specific population by weighting the transition
matrix elements with the occupation probability of, rather than just
averaging over the decaying substates.
The population of the various nuclear substates
is incorporated in the formalism by introducing
the density matrix of the excited quantum system or, in the case
of rotational symmetry of the problem, by a set of statistical tensors
which obey the
same transformation law as the spherical harmonics. This concept enables us
to treat the polarization or alignment of excited nuclei appropriately.
First calculations of the angular correlation of electrons and positrons
emitted in internal pair conversion taking into account the alignment
of nuclei were accomplished by Goldring, Rose and Warburton
\cite{goldring,rose:63,warburton}. These calculations were performed within
Born approximation, neglecting the influence of the nuclear charge on the
outgoing electron and positron. But for
internal pair conversion (IPC) of highly charged nuclei the
Born approximation is not justified as can be verified by the corresponding
positron spectra \cite{schlueter:78,schlueter:81,soff:81}.
Therefore we reconsider in the following the internal pair conversion
of heavy nuclei which
are aligned, e.g., by Coulomb excitation or transfer reactions.
We determine the angular correlation
of the emitted electron and positron with respect
to a reference axis in space. As
already known for the angular correlation of $\gamma$ rays, the problem will
be simplified if we choose a coordinate system in which the density matrix is
diagonal. The statistical tensors depend as well on the choice of the
coordinate system. If the entries of the statistical tensors
are given in a specific coordinate system, we are able
to calculate the angular correlation with respect to the $z$ axis of
this system.
The occupation probabilities of the magnetic substates
caused by Coulomb excitation can be calculated with, e.g.,
the COULEX code of K.~Alder and A.~Winther \cite{alder:winther:1}.
However, one should take into account the change of
population by electromagnetic transitions from
higher lying states. Special attention should be paid
to a proper choice of the coordinate system when dealing with the COULEX code
\cite{alder:winther:1,alder:winther:2}.
For pure Coulomb excitation we will assume the $z$ axis to point
along the asymptotic target recoil axis. With respect to this axis
the excited nuclei may exhibit prolate or oblate alignment.
\section{Density matrix and statistical tensors}
The density matrix -- and for spherical symmetry the set of statistical
tensors -- is the appropriate tool for including statistical
properties such as
occupation probabilities of quantum mechanical states
into the calculations.
Here we briefly summarize the essential properties of the density matrix
and subsequently turn
to the concept of the statistical tensors, which obey the same
transformation law as irreducible tensors. For the density matrix
$\rho_{M_i M'_i}(J_i)$ of dimension $(2J_i+1)\times(2J_i+1)$ we note:
\begin{enumerate}
\item
The density matrix is hermitian:
$$ \rho^{*}_{M'_i M_i}( J_i ) = \rho_{M_i M'_i}( J_i ) \quad , $$
\item
The trace of the density matrix equals one:
$$
{{\rm Tr} \{ \rho \} } = 1 \Leftrightarrow
\sum\limits_{M_i} \rho_{M_i M_i}( J_i ) = 1 \quad , $$
\item
${\rm Tr} \{ \rho^2 \} \leq 1 \quad,\quad \mbox{and} \quad
{\rm Tr} \{ \rho^2 \} = 1 \Leftrightarrow
\mbox{the system is in a pure state.}$
\end{enumerate}
We define the statistical tensors $\hat\rho^{[n]}_{\nu}$ as irreducible tensors
of rank $n$ with $\nu=-n,\ldots,n$:
\begin{equation}
\hat\rho^{[n]}_{\nu}(J_i) = \sum_{M_i,M'_i} (-1)^{J_i-M_i'} \sqrt{2n+1}
\left( \begin{array}{ccc} J_i & J_i & n \\ M_i & M'_i & -\nu
\end{array}\right) \rho_{M_i M'_i}( J_i )
\quad .
\label{dichtetensor-hin}
\end{equation}
The argument $J_i$ reminds us that $n$ is related to the angular momentum
of the magnetic substates by $0 \leq n \leq 2J_i$.
From the normalization of the density matrix it follows
$\hat\rho^{[0]}_{0}(J_i) = 1/\sqrt{2J_i+1}$.
The density matrix has $(2J_i+1)^2$ independent components. To describe a
system by statistical tensors instead of the density matrix, we need
$2J_i+1$ density tensors of rank $n=0$ up to rank $n=2J_i$. Since
the density tensor of rank $n$ has $2n+1$ components we get again
$\sum_{n=0}^{2J_i}\left(\sum_{\nu=-n}^{n} 1 \right) = (2J_i+1)^2$ independent
components.
The statistical tensors transform under rotations according to
\begin{equation}
\hat\rho^{[n]}_{\nu}(J_i) = \sum\limits_{\nu'}
{\cal D}^{[n]*}_{\nu' \nu}( \vec\alpha ) \, \hat\rho^{[n]}_{\nu'}(J_i)
\end{equation}
where the Wigner rotation matrix of rank $n$ is denoted by ${\cal D}^{[n]}$
and the set of Euler angles by $\vec\alpha$. In defining the Euler angles
we follow Rose \cite{rose:57} and Eisenberg\&Greiner
\cite{eisenberg:greiner:1}.
For systems with rotational symmetry it is thus more advantageous to
employ the concept of the statistical
tensors when incorporating statistical statements concerning the system.
The components of the statistical tensors are changed under rotations
and so are the occupation numbers of the magnetic substates.
From the set of $2J_i+1$ statistical tensors we obtain the density matrix by
utilizing the relation:
\begin{equation}
\rho_{M_i M_i'}( J_i ) = (-1)^{J_i-M'_i} \sum\limits_{n,\nu} \sqrt{2n+1}
\left(\begin{array}{ccc} J_i & J_i & n \\ M_i' & -M_i & -\nu \end{array}\right)
\hat\rho^{[n]}_{\nu}(J_i)
\label{dichtetensor-rueck}
\quad .
\end{equation}
For certain symmetries of the system we can reduce the independent components
of the statistical tensors.
Here we list the consequences for the statistical tensors in three special
cases which will become relevant for us:
\begin{enumerate}
\item
In the case of {\em axial symmetry} the density matrix is diagonal, its
diagonal components are just the probabilities for the occupation of the
corresponding magnetic substates $\rho_{M_i M_i} = p_{M_i}$:
$$
\hat\rho^{[n]}_{\nu}(J_i) =
\delta_{0\nu}\sum\limits_{M_i} (-1)^{J_i-M_i} p_{M_i}
\left(\begin{array}{ccc} J_i & J_i & n \\ M_i & -M_i & 0
\end{array}\right)
\quad .
$$
One can always choose a basis such that the density matrix is diagonal,
but in general this will not be a basis of wave functions with good
angular momentum.
\item
In the case of {\em spherical symmetry}, there is no direction singled
out in space.
The density matrix is proportional to the identity matrix. The diagonal
elements are given by $\rho_{M_i M_i} = 1/(2J_i+1)$. All statistical tensors
vanish with exception of the tensor of rank $0$, i.e.,
$$
\hat\rho^{[n]}_{\nu}(J_i) = \delta_{0n}\,\delta_{0\nu}
\frac{1}{\sqrt{2J_i+1}}
\quad.
$$
\item
From Eq.~(\ref{dichtetensor-hin}) it can be shown that for alignment
of the nuclear states, defined by $ p_{M_i} = p_{-M_i} $,
the statistical tensors of odd rank vanish.
\end{enumerate}
\section{Angular correlation of $\gamma$-rays}
Before we enter into the calculations concerning the angular correlation
in internal pair conversion we summarize some results already known
for in-beam $\gamma$-ray spectroscopy.
This will help us to interpret the angular correlation pattern in the case
of internal pair conversion. The angular correlation of photons emitted
after Coulomb excitation is
essentially determined by the statistical tensors, i.e.,
by the occupation numbers of the magnetic substates of the decaying nucleus.
In choosing a reference axis for which the density matrix is diagonal, just
the 0th components of all statistical tensors survive and we obtain for the
transition probability the well-known relations:
\begin{equation}
\frac{{\rm d}P_\gamma}{{\rm d}\Omega}
= \frac{2\alpha\omega}{\sqrt{2J_i+1}} \left| V_\gamma^{(\tau)}(L) \right|^2
\sum\limits_{I\ {\rm even}} F_I( L \, L \, J_f \, J_i )
\; \hat\rho^{[I]}_{0}(J_i) \, P_I(\cos\vartheta)
\quad ,
\end{equation}
for a transition of parity $\tau={\rm E/M}$ and multipolarity $L$.
$V_\gamma^{(\tau)}(L)$ denotes the corresponding reduced matrix
element for the nuclear transition.
Here we employed the correlation coefficients
\cite{biedenharn,schwalm,wollersheim}
\begin{eqnarray}
F_I( L \, L \, J_f \, J_i ) & = & (-)^{J_f+J_i -1} \,
\sqrt{2I+1} \, \sqrt{ 2J_i+1}
\nonumber
\\
& \times &
(2L+1)
\left(\begin{array}{ccc} L & L & I \\ 1 & -1 & 0 \end{array}\right)
\left\{\begin{array}{ccc} L & L & I \\ J_i & J_i & J_f \end{array}\right\}
\label{ph-emission-20}
\; .
\end{eqnarray}
This results in an anisotropic emission of photons with respect to the
alignment axis. The number of minima of the angular distribution corresponds
to the multipolarity of the nuclear transition.
In the case of spherical symmetry, the photon emission is isotropic
\begin{equation}
\frac{{\rm d}P_\gamma}{{\rm d}\Omega} =
\frac{2\alpha\omega}{2J_i+1} \, \left| V_\gamma^{(\tau)}(L)\right|^2
\end{equation}
or integrated over the solid angle $\Omega$:
\begin{equation}
P_\gamma = \frac{8\pi\alpha\omega}{2J_i+1} \,
\left| V_\gamma^{(\tau)}(L)\right|^2
\quad .
\label{photon}
\end{equation}
\section{Transition probabilities for internal pair\protect\newline
conversion}
We turn now to the formulation of the triple correlation of the
electron and positron direction with reference to a symmetry axis,
which is taken as quantization axis.
For a statistical ensemble of nuclei we write the transition probability
for internal pair conversion,
\begin{eqnarray}
P_{e^{+}e^{-}}
& = &
2\pi
\sum\limits_{M_i , M_i' , M_f , \lambda , \lambda'}
\int {\rm d}^{3}p \int {\rm d}^{3}p' \, \delta(\omega - W' - W)
\nonumber
\\
& &
\times
U_{\rm pl}\; \rho_{M_i M_i'} \; U^{*}_{\rm pl}
\quad .
\label{p}
\end{eqnarray}
where the density matrix $\rho_{M_i M'_i}$ represents the occupation of the
magnetic substates $|J_i M_i\rangle$.
Here we assumed a nuclear transition from a initial state $|J_i M_i\rangle$
to the final state $|J_f M_f\rangle$ where the initial state is populated
according to the density matrix $\rho_{M_i M'_i}$.
Since we do not require the density matrix to be diagonal the
summation extends over both, $M_i$ and $M'_i$. The $\delta$ function
ensures energy conservation: The transition energy $\omega$ is transfered
to the electron (total energy $W'$) and to the positron (total energy $W$).
The summation is taken over the spins and the momenta of the outgoing leptons.
The matrix element for internal pair conversion is written in lowest
order of $\alpha$ in the retarded form:
\begin{equation}
U_{{\rm pl}} = -\alpha\int {\rm d} V_{\rm n} \int {\rm d} V_{\rm e}
\left(
\rho_{\rm n}(\vec r_{\rm n}) \, \rho_{\rm e}(\vec r_{\rm e})
-\vec j_{\rm n}(\vec r_{\rm n}) \cdot \vec j_{\rm e}(\vec r_{\rm e})
\right) \, \frac{e^{{\rm i}\omega| \vec r_{\rm n} - \vec r_{\rm e} |}}
{\left| \vec r_{\rm n} - \vec r_{\rm e}\right|}
\quad,
\label{u}
\end{equation}
$\vec r_{\rm e}$ being the electronic coordinate, $\vec r_{\rm n}$ the
nuclear coordinate.
Since we neglect in our work the penetration of the electron wave functions
we do not have to specify the nuclear transition charge and current densities
$\rho_{\rm n}$ and $\vec j_{\rm n}$. The electronic transition charge and
current densities read
\begin{equation}
\rho_{\rm e} = \psi_{\rm f}^\dagger \, \psi_{\rm i}
\quad , \qquad
\vec j_{\rm e} = \psi_{\rm f}^\dagger \,\vec\alpha\, \psi_{\rm i}
\end{equation}
($\vec\alpha$ is the 3-vector of the spatial Dirac matrices in the standard
representation). These expressions are evaluated utilizing the scattering
solutions, see Eqs.~(\ref{eswelle},\ref{pswelle}) in the appendix,
for the electron and positron wave
function in order to define the emission direction and thus an opening angle.
Inserting the spherical wave expansion of
these wave functions results in a decomposition of the matrix element,
Eq.~(\ref{u}),
\begin{eqnarray}
U_{\rm pl} = \sum\limits_{\kappa' , \mu'} \sum\limits_{\kappa , \mu}
a^{(-) *}_{\kappa' \mu'} \; b^{(+)}_{\kappa \mu} \;
U_{\kappa'\mu' \kappa\mu}
\quad .
\end{eqnarray}
$U_{ \kappa'\mu' \kappa\mu }$
denotes the transition matrix element which has the same structure
as $U_{\rm pl}$, but is evaluated
using the spherical spinor solutions of the Dirac equation, Eqs.~(\ref{ewelle}).
This matrix element was calculated in \cite{schlueter:81}. Here we cite the
result:
\begin{eqnarray}
U_{ \kappa'\mu' \kappa\mu }
& = &
4\pi{\rm i}\alpha\omega \, (-1)^{J_f-M_f}
\left(\begin{array}{ccc} J_f & L & J_i \\ -M_f & M & M_i
\end{array}\right)
V_{\gamma}^{(\tau)}(L)
\nonumber
\\
& &
\times (-1)^{j'-\mu'}
\left(\begin{array}{ccc} j' & L & j \\ -\mu' & M & \mu
\end{array}\right)
M_{\kappa' \kappa}^{(\tau)}(L)
\quad .
\label{usph}
\end{eqnarray}
$V_{\gamma}^{(\tau)}(L)$ is just the reduced nuclear matrix element
of Eq.~(\ref{photon}) and
\begin{eqnarray}
M^{(\tau)}_{\kappa' \kappa}(L)
& = &
-{\rm i} \, (-1)^{j'+\frac{1}{2}} \,
\frac{\sqrt{2j+1}\,\sqrt{2j'+1}\,\sqrt{2L+1}}{4\pi\sqrt{L(L+1)}}
\nonumber\\
& &
\times \left(\begin{array}{ccc} j & j' & L \\ -1/2 & 1/2 & 0
\end{array}\right)
R^{(\tau)}_{\kappa' \kappa}
\label{ematrix}
\end{eqnarray}
with the parity selection rule
\begin{equation}
l + l' + L + \lambda(\tau) = 0 \ \mbox{mod}\ 2 \quad,
\quad
\left\{ \begin{array}{ccl} \lambda=0 & \mbox{for} & \tau = {\rm el} \\
\lambda=1 & \mbox{for} & \tau = {\rm magn}
\end{array}\right.
\end{equation}
$R^{(\tau)}_{\kappa' \kappa}$ contains the integration over the radial
electron wave functions and will be defined later.
Inserting this matrix element into the pair conversion probability,
Eq.~({\ref p}), yields
\begin{eqnarray}
P_{e^{+}e^{-}}
& = &
2\pi
\sum\limits_{M_i , M_i' , M_f} \rho^{[J_i]}_{M_i M_i'}
\sum\limits_{\lambda , \lambda'}
\int{\rm d} W\,{\rm d}\Omega \, \int{\rm d} W'\,{\rm d}\Omega \,
\delta(\omega - W' - W)
\nonumber
\\
& &
\times
\sum\limits_{\kappa' , \mu'} \sum\limits_{\kappa , \mu}
\sum\limits_{\bar\kappa' , \bar\mu'} \sum\limits_{\bar\kappa , \bar\mu}
A_{\kappa' \mu' ; \bar\kappa' \bar\mu'} \,
B_{\kappa \mu ; \bar\kappa \bar\mu } \,
U_{\kappa' \mu' \kappa \mu } \,
U^{*}_{\bar\kappa' \bar\mu' \bar\kappa \bar\mu }
\quad .
\label{Pep}
\end{eqnarray}
where we abbreviated
\begin{equation}
A_{\kappa' \mu' ; \bar\kappa' \bar\mu'} = W'\,p'\sum\limits_{\lambda'}
a^{(-) *}_{\kappa' \mu'} \, a^{(-)}_{\bar\kappa' \bar\mu'}
\label{A}
\end{equation}
and
\begin{equation}
B_{ \kappa \mu ; \bar\kappa \bar\mu } = W\,p\sum\limits_{\lambda}
b^{(+)}_{\kappa \mu} \, b^{(+) *}_{\bar\kappa \bar\mu}
\quad .
\label{B}
\end{equation}
From Eq.~(\ref{Pep}) we obtain the differential pair conversion probability
with respect to the {\em kinetic positron energy} $E = W - m$ and the solid
angles of both electron, $\Omega'$, and positron, $\Omega$,
\begin{equation}
P_{e^+e^-} = \int\limits_0^{\omega-2m} {\rm d}E \int{\rm d}\Omega
\int{\rm d}\Omega' \, \frac{{\rm d}^3 P_{e^+ e^-}}{{\rm d}E \,
{\rm d}\Omega \, {\rm d}\Omega'}
\quad .
\label{int}
\end{equation}
The integration over the electron energy $W'$ is trivially performed because
of the $\delta$ function. From this relation it is obvious that we may proceed
from the solid angles $\Omega$ to $\tilde\Omega$ by choosing another
reference axis in space. The integrand in Eq.~(\ref{int}) is invariant
under rotations since the Jacobian of this transformation equals 1. The
integrand should thus be represented by a series of triple correlation functions
which are defined in Eq.~(\ref{tripelkorrelationsfunktion}).
Inserting the explicit expressions of the coefficients $A$ and $B$,
Eqs.~(\ref{koeff-a},\ref{koeff-b}) leads to the following expression
for the differential pair conversion probability
\begin{eqnarray}
\lefteqn{
\frac{{\rm d}^3 P_{e^+ e^-}}{{\rm d}E\,{\rm d}\Omega\,{\rm d}\Omega'}
=
8(\pi\alpha\omega)^2 \, |V_\gamma^{(\tau)}(L)|^2
\sum\limits_{M , \bar M} \sum\limits_{M_i , M_i'} \sum\limits_{M_f}
\rho^{[J_i]}_{M_i M_i'}
}
\nonumber
\\
& &
\times
\left(\begin{array}{ccc} J_f & L & J_i \\ -M_f & M & M_i
\end{array}\right)
\left(\begin{array}{ccc} J_f & L & J_i \\ -M_f & \bar M & M_i'
\end{array}\right)
\nonumber
\\
& &
\times
\sum\limits_{\kappa , \kappa' , \bar\kappa , \bar\kappa'}
(-1)^{j'+\bar j'} \, M^{(\tau)}_{\kappa' \kappa}(L) \,
M^{(\tau) *}_{\bar\kappa' \bar\kappa}(L)
\sqrt{2\bar j'+1} \, \sqrt{2j'+1}
\nonumber
\\
& &
\times
\sqrt{2j+1} \, \sqrt{2\bar j+1} \,
\nonumber
\\
& &
\times
\exp({\rm i}[(\delta'(W',\kappa')-\bar\delta'(W',\bar\kappa')
+\delta(-W,\kappa)-\bar\delta(-W,\bar\kappa)])
\nonumber
\\
& &
\times \sum\limits_{I' , I , \alpha , \alpha'}
\sqrt{2I'+1} \, \sqrt{2I+1} \,
Y_{I'\alpha'}(\Omega_{p'}) \, Y_{I\alpha}(\Omega_{p})
\nonumber
\\
& &
\times
\left(\begin{array}{ccc} j' & \bar j' & I' \\ 1/2 & -1/2 & 0
\end{array}\right)
\left(\begin{array}{ccc} j & \bar j & I \\ 1/2 & -1/2 & 0
\end{array}\right)
\nonumber
\\
& &
\times
\sum\limits_{\mu , \mu' , \bar\mu , \bar\mu'}
(-1)^{\bar\mu-\bar\mu'+1}
\left(\begin{array}{ccc} \bar j' & L & \bar j \\ -\bar\mu' & \bar M & \bar\mu
\end{array}\right)
\nonumber
\\
& &
\times
\left(\begin{array}{ccc} j' & L & j \\ -\mu' & M & \mu
\end{array}\right)
\left(\begin{array}{ccc} \bar j' & j' & I' \\ -\bar\mu' & \mu' & -\alpha'
\end{array}\right)
\left(\begin{array}{ccc} \bar j & j & I \\ -\bar\mu & \mu & \alpha
\end{array}\right)
\quad .
\end{eqnarray}
Here we inserted Eq.~(\ref{usph}).
Introducing the statistical tensors we are left with
\begin{eqnarray}
\frac{{\rm d}^3 P_{e^+ e^-}}{{\rm d}E\,{\rm d}\Omega\,{\rm d}\Omega'}
& = &
2(4\pi\alpha\omega)^2 \, |V_\gamma^{(\tau)}(L)|^2 \,
(-1)^{J_f-J_i+L+1}
\nonumber
\\
& &
\times
\sum\limits_{n , \nu} \sqrt{2n+1} (-1)^{\nu} \hat\rho^{[n]}_{\nu}(J_i)
\left\{\begin{array}{ccc} L & L & n \\ J_i & J_i & J_f \end{array}\right\}
\nonumber
\\
& &
\times \sum\limits_{I , I'}
\sqrt{2I+1} \, \sqrt{2I'+1} \, (-1)^{I'} \,
\nonumber
\\
& &
\times
\sum\limits_{\alpha , \alpha'}
Y_{I'\alpha'}(\Omega_{p'}) \, Y_{I\alpha}(\Omega_{p})
\left(\begin{array}{ccc} I & I' & n \\ \alpha & \alpha' & -\nu
\end{array}\right)
\nonumber
\\
& &
\times
\sum\limits_{\kappa , \kappa' , \bar\kappa , \bar\kappa'}
(-1)^{\bar j+\bar j'} \,
\sqrt{|\kappa \, \kappa' \, \bar\kappa \, \bar\kappa'|}
\left\{\begin{array}{ccc} \bar j & \bar j' & L \\
j & j' & L \\
I & I' & n \end{array}\right\}
M^{(\tau)}_{\kappa' \kappa}(L) \, M^{(\tau) *}_{\bar\kappa' \bar\kappa}(L)
\nonumber
\\
& &
\times
\exp({\rm i}[\delta'(E',\kappa')-\bar\delta'(E',\bar\kappa')+
\delta(-E,\kappa)-\bar\delta(-E,\bar\kappa)])
\nonumber
\\
& &
\times
\left(\begin{array}{ccc} j' & \bar j' & I' \\ 1/2 & -1/2 & 0
\end{array}\right)
\left(\begin{array}{ccc} j & \bar j & I \\ 1/2 & -1/2 & 0
\end{array}\right)
\quad.
\end{eqnarray}
This is the most general form for the pair conversion probability. Now
we assume that we are dealing with internal pair conversion of aligned
nuclei ($\nu=0$).
We may choose an appropriate coordinate system by transformation
of the spherical harmonics:
\begin{eqnarray}
Y_{I\alpha}(\Omega_{p})
& = &
\sum\limits_{\beta}
\exp({\rm i}\alpha\phi) \, d^{[I]}_{\alpha\beta}(\vartheta) \,
\exp({\rm i}\beta\delta) \, Y_{I\beta}( 0, 0 )
\nonumber\\
& = &
\sqrt{\frac{2I+1}{4\pi}} \, \exp({\rm i}\alpha\phi) \,
d^{[I]}_{\alpha 0}(\vartheta)
\nonumber
\\
Y_{I'\,-\alpha}(\Omega_{p'})
& = &
\sum\limits_{\beta'} \exp(-{\rm i}\alpha\phi) \,
d^{[I']}_{-\alpha\beta'}(\vartheta) \, \exp({\rm i}\beta'\delta)
\, Y_{I'\beta'}(\Theta, 0)
\end{eqnarray}
Here $\Theta$ denotes the opening angle of the electron-positron pair,
$\vartheta$ is the polar angle of the positron with respect to the symmetry
axis, and the dihedral angle $\delta$ indicates the rotation of the
electron-positron plane around the positron axis (Fig.~2b). Please note,
that the convention of \cite{goldring,rose:63,warburton} differs in the
definition of the angles from the one employed here.
This enables us to define the triple correlation function by
\begin{eqnarray}
\lefteqn{P_{II'n}(\vartheta,\Theta,\delta) =
\sum\limits_{\alpha}
\left(\begin{array}{ccc} I & I' & n \\ \alpha & -\alpha & 0
\end{array}\right)
Y_{I\alpha}(\Omega_{p}) \, Y_{I'\,-\alpha'}(\Omega_{p'})
}
\nonumber
\\
& = &
\frac{\sqrt{2I+1} \, \sqrt{2I'+1}}{4\pi}
\sum\limits_{\beta'} (-1)^{\beta'}
\left(\begin{array}{ccc} I & I' & n \\ 0 & \beta & \beta'
\end{array}\right)
d^{[n]}_{\beta' 0}(\vartheta) \, d^{[I']}_{\beta' 0}(\Theta) \,
\exp({\rm i}\beta'\delta)
\quad .
\nonumber
\\
& &
\label{tripelkorrelationsfunktion}
\end{eqnarray}
Our triple correlation function is related to the one introduced
by Biedenharn \cite{biedenharn} by a factor
$4\pi{\rm i}^{-I-I'-n}/((2I+1)(2I'+1))^{1/2}$.
The pair conversion probability is normalized by the probability for
$\gamma$ emission, Eq.~(\ref{photon}), which yields the pair conversion
coefficient:
\begin{eqnarray}
\lefteqn{
\frac{{\rm d}^4 \beta}{{\rm d}E\,{\rm d}\cos\Theta\,{\rm d}\cos\vartheta
\,{\rm d}\delta} =
\frac{2\alpha\omega (2L+1)}{L(L+1)} \, (2J_i+1) \,
(-1)^{J_f-J_i+L+1} }
\nonumber
\\
& &
\times
\sum\limits_{n} \sqrt{2n+1} \, \hat\rho^{[n]}_{0}(J_i) \,
\left\{\begin{array}{ccc} L & L & n \\ J_i & J_i & J_f
\end{array}\right\}
\nonumber
\\
& &
\times \sum\limits_{I , I'} (2I+1) (2I'+1) (-1)^{I'}
\nonumber
\\
& &
\times \sum\limits_{\beta'} (-1)^{\beta'}
\left(\begin{array}{ccc} I & I' & n \\ 0 & \beta' & -\beta'
\end{array}\right)
d^{[n]}_{\beta' 0}(\vartheta) \, d^{[I']}{\beta' 0}(\Theta) \,
\exp({\rm i}\beta'\delta)
\nonumber
\\
& &
\times \sum\limits_{\kappa , \kappa' , \bar\kappa , \bar\kappa'}
(-1)^{j+\bar j'} |\kappa \, \kappa' \, \bar\kappa \, \bar\kappa'|
\left\{\begin{array}{ccc} \bar j & \bar j' & L \\ j & j' & L \\
I & I' & n \end{array}\right\}
R^{(\tau)}_{\kappa' \kappa}(L) \, R^{(\tau) *}_{\bar\kappa'\bar\kappa}(L)
\nonumber
\\
& &
\times \exp({\rm i}[\delta'(W',\kappa')-\bar\delta'(W',\bar\kappa')+
\delta(-W,\kappa)-\bar\delta(-W,\bar\kappa)]) \,
\nonumber
\\
& &
\times
\left(\begin{array}{ccc} j' & \bar j' & I' \\ 1/2 & -1/2 & 0
\end{array}\right)
\left(\begin{array}{ccc} j & \bar j & I \\ 1/2 & -1/2 & 0
\end{array}\right)
\nonumber
\\
& &
\times
\left(\begin{array}{ccc} j & j' & L \\ 1/2 & -1/2 & 0
\end{array}\right)
\left(\begin{array}{ccc} \bar j & \bar j' & L \\ 1/2 & -1/2 & 0
\end{array}\right)
\quad .
\label{ergebnis}
\end{eqnarray}
Here we inserted the explicit expressions for the electronic matrix
elements, Eq.~(\ref{ematrix}). Integration over the azimuthal angle is
trivially performed resulting in an additional factor of $2\pi$.
The radial matrix elements read for electric pair conversion (parity
$(-)^{L}$)
\begin{equation}
R^{({\rm e})}_{\kappa' \kappa} =
L ( R_1 + R_2 + R_3 - R_4 ) + ( \kappa - \kappa' ) ( R_3 + R_4 )
\label{erade}
\end{equation}
and for magnetic pair conversion (parity $(-)^{L+1}$)
\begin{equation}
R^{({\rm m})}_{\kappa' \kappa} =
(\kappa + \kappa') ( R_5 + R_6 )
\quad .
\label{eradm}
\end{equation}
The radial integrals introduced in these equations are taken over products of
the radial electron wave functions (\ref{radiale}) and the Hankel functions
of first kind, $h^{(1)}_L(\omega r)$:
\begin{eqnarray}
R_1 & = & \int\limits_0^\infty {\rm d} r \, r^2 \, g_{W',\kappa'}(r) \,
g_{-W,\kappa}(r)
\,
h^{(1)}_L(\omega r) \quad , \nonumber\\
R_2 & = & \int\limits_0^\infty {\rm d} r \, r^2 \, f_{W',\kappa'}(r) \,
f_{-W,\kappa}(r)
\,
h^{(1)}_L(\omega r) \quad , \nonumber\\
R_3 & = & \int\limits_0^\infty {\rm d} r \, r^2 \, g_{W',\kappa'}(r) \,
f_{-W,\kappa}(r)
\,
h^{(1)}_{L-1}(\omega r) \quad , \nonumber\\
R_4 & = & \int\limits_0^\infty {\rm d} r \, r^2 \, f_{W',\kappa'}(r) \,
g_{-W,\kappa}(r)
\,
h^{(1)}_{L-1}(\omega r) \quad , \nonumber\\
R_5 & = & \int\limits_0^\infty {\rm d} r \, r^2 \, g_{W',\kappa'}(r) \,
f_{-W,\kappa}(r)
\,
h^{(1)}_L(\omega r) \quad , \nonumber\\
R_6 & = & \int\limits_0^\infty {\rm d} r \, r^2 \, f_{W',\kappa'}(r) \,
g_{-W,\kappa}(r)
\,
h^{(1)}_L(\omega r)
\label{radialintegrale}
\quad.
\end{eqnarray}
In the case of a point-like nucleus these integrals can be rewritten in
terms of $F_2$ functions \cite{schlueter:78} which can be evaluated numerically.
For the representation of the nucleus as a homogeneously charged
sphere the radial integrals are computed using a Gauss-Chebyshev quadrature
\cite{chebyshev}.
The Whittaker functions which occur in the expressions for the electron
wave functions are computed with the COULCC code of \cite{barnett}.
Since the integrands are oscillating functions
it is advantageous to deform the
integration contour in the complex plane
in such a way that it runs along the imaginary axis
\cite{schlueter:81}.
Since the electron wave functions have the asymptotic behaviour
$\exp({\rm i} p r)$ while the Hankel functions behave like
$\exp({\rm i} \omega r)$, where $\omega=W+W'$, the integrand for
large $r$ assumes the form:
\begin{equation}
\exp({\rm i}[-p-p'+W+W']r)
\quad .
\end{equation}
For $r$ complex with the imaginary part going to infinity, our procedure thus
guarantees that the integrand falls off quite fast. In most cases
at $r=20000$ fm the integrand is smaller than $10^{-5}$ of its maximum value.
We want to consider the angular correlation for two special cases:
I.
If we integrate Eq.~(\ref{ergebnis})
over the positron polar angle $\vartheta$ and the dihedral
angle $\delta$ the remaining function depends only on the opening angle
$\Theta$ of the electron-positron pair. In this case
only the $n=0$ contribution survives.
We get the opening angle
distribution as a series of Legendre polynomials which was already calculated
in \cite{hofmann:90}:
\begin{equation}
\frac{{\rm d}^2\beta}{{\rm d}E\,{\rm d}\cos\Theta} =
\sum\limits_{I} a_I \, P_I(\cos\Theta)
\end{equation}
The expansion coefficients are given by
\begin{eqnarray}
a_I & = &
\frac{8\pi\alpha\omega}{L(L+1)} \, (-)^{L+I+1} (2I+1)
\sum\limits_{\kappa,\kappa',\bar\kappa,\bar\kappa'} \left|
\kappa\,\kappa'\,\bar\kappa\,\bar\kappa'\right|
R^{(\tau)}_{\kappa'\kappa}(L) \, R^{(\tau)*}_{\bar\kappa'\bar\kappa}
\nonumber\\
& \times &
\exp({\rm i}[\delta'(W',\kappa') - \bar\delta'(W',\bar\kappa')
+\delta(-W,\kappa) - \bar\delta(-W,\bar\kappa)])
\nonumber\\
& \times &
\left(\begin{array}{ccc} j' & \bar j' & I \\ 1/2 & -1/2 & 0 \end{array}\right)
\left(\begin{array}{ccc} j & \bar j & I \\ 1/2 & -1/2 & 0 \end{array}\right)
\nonumber\\
& \times &
\left(\begin{array}{ccc} j & j' & L \\ 1/2 & -1/2 & 0 \end{array}\right)
\left(\begin{array}{ccc} \bar j & \bar j' & L \\ 1/2 & -1/2 & 0
\end{array}\right)
\left\{\begin{array}{ccc} \bar j & \bar j' & L \\ j' & j & L \end{array}
\right\}
\quad .
\end{eqnarray}
They have to be evaluated numerically.
The same result is achieved if one assumes that the initial nuclear
substates are equally populated.
At this point we apologize for
giving an incorrect expression for the opening angle distribution in
\cite{hofmann:90} which was caused by employing the wrong set of
scattering solutions.
This error resulted in the wrong sign
of the scattering phase shifts of the positron. The opening angle
distribution showed the right qualitative behaviour but wrong conversion
probabilities. The statement, that the maximum of the distribution shifts
from $0^\circ$ to $180^\circ$, if one considers
overcritical nuclear charges ($Z\geq 173$), remains unchanged. This
error appeared also in the expression for the electric monopole (E0)
conversion. One should reverse the sign of the scattering phase shifts
of the positron. The pair conversion coefficient for
the electric monopole conversion reads:
\begin{equation}
\frac{{\rm d}\eta}{{\rm d}E \, {\rm d}\cos\Theta}
= \frac{1}{2} \, \frac{{\rm d}\eta}{{\rm d}E}
( 1 + \varepsilon \, \cos\Theta )
\label{maximum}
\end{equation}
where ${\rm d}\eta / {\rm d}E$ is the differential
pair conversion coefficient \cite{soff:81}
---which remains unchanged--- and $\varepsilon$ is the
corrected anisotropy coefficient:
\begin{equation}
\varepsilon = 2 \, \frac{C_{-1}\,C_{+1}}{C_{-1}^2 + C_{+1}^2} \,
\cos(\Delta_{+1 \, -1})
\end{equation}
with
\begin{equation}
\Delta_{\kappa\kappa'} =
\delta'(W',\kappa)-\delta'(W',\kappa')+
\delta(-W,\kappa)-\delta(-W,\kappa')
\quad.
\end{equation}
$C_{+1}$, $C_{-1}$ are defined by
\begin{equation}
C_\kappa = \left\{ \begin{array}{ccc}
\displaystyle\lim_{r\to 0} \frac{f_{-W,\kappa}(r) \, f'_{W',\kappa}(r)}
{r^{2j-1}} & \mbox{for} & \kappa>0
\\[3mm]
\displaystyle\lim_{r\to 0} \frac{g_{-W,\kappa}(r) \, f'_{W',\kappa}(r)}
{r^{2j-1}} & \mbox{for} & \kappa<0
\end{array}\right .
\end{equation}
where $f$ ($f'$) and $g$ ($g'$) are the radial wave functions of the
Dirac spinor of the positron (electron). In numerical calculations these
constants are evaluated at the nuclear radius.
$\delta$ and $\delta'$ are the corresponding Coulomb phase shifts
for an extended nucleus \cite{mueller:rafelski:greiner:3}.
II.
If we integrate Eq.~(\ref{ergebnis}) over the opening angle $\Theta$ of the electron-positron
pair and the dihedral angle $\delta$ we end up with
\begin{equation}
\frac{{\rm d}^2\beta}{{\rm d}E\,{\rm d}\cos\vartheta} =
\sum\limits_n b_n \, P_n(\cos\vartheta)
\end{equation}
where the coefficients read
\begin{eqnarray}
b_n & = &
4\pi\alpha\omega \, \frac{(2L+1)(2J_i+1)}{L(L+1)} \, (-)^{J_f - J_i + 1}
\sqrt{2n+1} \, \hat\rho^{[n]}_0(J_i)
\left\{\begin{array}{ccc} L & L & n \\ J_i & J_i & J_f \end{array}\right\}
\nonumber\\
& \times &
\sum\limits_{\kappa,\kappa',\bar\kappa} (-)^{j+\bar j + j'+1/2}
\left|\kappa\,\kappa'\,\bar\kappa\right|
\left\{\begin{array}{ccc} L & L & n \\ j & \bar j & j' \end{array}\right\}
R^{(\tau)}_{\kappa'\kappa}(L) \, R^{\tau *}_{\kappa'\bar\kappa}(L)
\nonumber\\
& \times &
\exp({\rm i}[\delta(-W,\kappa) - \bar\delta(-W,\bar\kappa)])
\nonumber\\
& \times &
\left(\begin{array}{ccc} j & \bar j & n \\ 1/2 & -1/2 & 0 \end{array}\right)
\left(\begin{array}{ccc} j & j' & L \\ 1/2 & -1/2 & 0 \end{array}\right)
\left(\begin{array}{ccc} \bar j & j' & L \\ 1/2 & -1/2 & 0 \end{array}\right)
\quad .
\end{eqnarray}
This corresponds to the experimental setup where one is just interested
in the angular distribution of the positron emitted in internal pair
conversion of an aligned or polarized nucleus.
\section{Results}
In the following we will discuss the characteristic properties of IPC
angular distributions using a few representative results. Since we
are interested in Coulomb effects they all refer to a uranium-like
nucleus ($Z=92$). The chosen energies and multipolarities are
generic and are not intended to represent particular nuclear
transitions known from experiment.
The opening angle distribution of electron and positron emitted by
IPC depicts for electric transitions the typical pattern:
it has its maximum at $\Theta=0^\circ$ and
its minimum for $\Theta=180^\circ$. For magnetic transitions in
heavy nuclei, however, the situation might be different.
Fig.~1a and 1b depict the
opening angle distribution for an E1 and a M1 transition of a uranium-like
nucleus as a result of our distorted wave Born approximation (DWBA)
in comparison with the Born approximation (BA). This demonstrates
how the angular correlation of the electron-positron pairs is
influenced by the strong Coulomb field of the nucleus.
In Fig.~1b we plotted also the opening-angle distribution
for the M1 transition taking into account the finite extension of the
nucleus under consideration. This verifies that the magnetic transitions
---and especially the M1 transition--- are very sensitive to the charge
distribution of the nucleus \cite{schlueter:81}.
For the E1 transition in Fig.~1a, on the other hand, the effect of the finite
nuclear size amounts to less than 0.1\,\%.
In the following we discuss
the triple angular correlation of electron and positron
for IPC of aligned nuclei. We take the symmetry axis as quantization axis
as in Eq.~(\ref{ergebnis}). The opening angle $\Theta$ of electron
and positron, the polar angle $\vartheta$ of the positron and the dihedral
angle $\delta$ form a complete set of angles to fix the emission directions
of electron and positron with respect to the symmetry axis.
The angles which describe the directions of the emitted leptons are displayed
in Fig.~2b. Note that our
choice of the coordinate system is different from that introduced
in the Born approximation calculations of
\cite{goldring,rose:63,warburton} in which $\vartheta$ denotes the polar
angle of the intermediate photon. However, since the Coulomb field
disturbs the momentum balance
we cannot determine the momentum of the intermediate
photon from the momenta of the outgoing leptons, which
would be necessary to
calculate the photon polar angle.
Depending on the experimental setup and reactions
various coordinate systems may be established in which the
statistical tensors are determined. Here we concentrate on
the Coulomb excitation of heavy ions in collisions
with beam energies at or below the
Coulomb barrier. In this case one usually chooses a coordinate system,
where the $z$ axis is pointing along the apex line
of the scattering hyperbola towards the projectile
and the $x$ axis is perpendicular to the scattering plane
(Fig.~2a). The $y$ axis
is then chosen such that the $y$ component of the projectile velocity
is positive \cite{alder:winther:1,alder:winther:2,broglia}.
In the sudden approximation it can be shown
that the nuclear states are excited with
a population of the magnetic substates reaching a maximum around
$M_i=0$, i.e.,
the nucleus is aligned in the plane perpendicular to the $z$ axis
(asymptotic recoil direction of the target)
\cite{wollersheim}. This is called oblate
alignment. Taking into account the de-excitation of the nucleus
by $\gamma$ cascades starting from high-spin the oblate alignment
changes into a prolate alignment with respect to the $z$ axis for the
low-spin states.
If the collision energy is increased the nuclear alignment changes
to a polarization with respect to a reference axis perpendicular to the
scattering plane \cite{alder:winther:2,broglia}. Classically
this corresponds to the situation where the drag caused by
surface friction puts the nuclei into a spinning motion.
After having chosen a coordinate system and having determined the degree
of alignment or polarization for the Coulomb excited nuclei
---the corresponding statistical tensors can be calculated with, e.g.,
the COULEX code of \cite{alder:winther:1}--- one can
employ Eq.~(\ref{ergebnis}) to determine the angular distribution of the
electron-positron pairs emitted by internal pair conversion of these
nuclei.
We plot in Fig.~3 the spatial correlation of the
electron-positron
pairs with respect to the reference axis assuming oblate alignment
of a uranium-like
nucleus. From the spectrum of the emitted pairs \cite{schlueter:81,schlueter:79}
we know that for large-$Z$ nuclei the
pair emission probability increases towards the maximum positron energy. Thus
the angular correlations are plotted for a case where nearly
the full transition energy (minus the electron rest mass)
is transferred to the positron. One recognizes a strong dependence
of the pair conversion probability on the polar angle of the positron with
respect to the reference axis. This behaviour resembles the anisotropic
emission of the intermediary photon
\cite{wollersheim}. The angular distribution
depends weakly on the dihedral angle $\delta$ of the electron-positron pair
(Fig.~4).
For transitions between nuclear states of high angular momentum the
opening angle distribution does not change drastically when the positron
polar angle is varied.
In order to elucidate the influence of the statistical tensors,
i.e., the occupation of the initial nuclear state on
the angular correlation of the emitted electron-positron pairs,
we present the angular distribution with respect to the polar
angle of the emitted positron.
Fig.~5 shows the polar angle
distribution assuming E1, E2 and E3 transitions to the $0^+$
ground state of nuclei which exhibit oblate
alignment. Fig.~6 displays the polar angle distribution
for a E1 transition to the $0^+$ ground state of a nucleus
for oblate and prolate alignment and for
polarization.
\section{Conclusion}
Angular correlations are very sensitive to the underlying process.
They may reveal a plethora of signatures for nuclear transition which
allow for an identification of the transition as well as for the study of the
properties of excited nuclei. Especially for large-$Z$ nuclei one cannot
rely on the validity of the Born approximation which becomes exact in
the limit $Z\rightarrow 0$. One rather has to perform the calculations
with the relativistic scattering wave functions for both electron and positron.
These wave functions take the
Coulomb distortion caused by the nuclear charge into account.
For magnetic transitions of large-$Z$ nuclei one has also to account for
the finite-size effects. Especially M1 transitions are very
sensitive on the extension of the nucleus. In order to study magnetic
IPC we approximated the nucleus undergoing the transition
by a homogeneously charged sphere.
The angular correlation of electron-positron pairs with respect to a
given axis in space depend on the statistical tensors which reflect the
population of the nuclear magnetic substates. The conversion probability
changes drastically when either the opening angle of the pair or the
polar angle of the positron is varied while the dependence on the
dihedral angle is rather weak.
Our calculations allow to make quantitative predictions of this
behaviour which qualitatively might have been anticipated from
the $\gamma$-ray spectroscopy performed in heavy-ion collisions.
Furthermore, the measurement
of the spatial correlation of electron-positron pairs can be
employed to obtain additional information about the nuclear transition.
E.g., not only the multipolarity but also the parity of the nuclear
transition can be measured in this way.
\\[1cm]
{\bf Acknowledgement:}
This work has been supported by the BMBF, by the
Deut\-sche For\-schungs\-gemein\-schaft (DFG), by GSI (Darmstadt),
and by the REHE programme of the European Science Foundation (ESF).
\newpage
|
\section{Introduction}
All the experimentally known fermions transform non-trivially
under the gauge group $SU(3)\times SU(2)\times U(1)$ of the standard model (SM).
However there are experimental
hints in the neutrino sector which suggest the existence of $SU(3)\times SU(2)\times U(1)$
- singlet fermions mixing appreciably with the known neutrinos.
These hints come from
(a) the deficits in the solar \cite{solar} and atmospheric \cite{atm}
neutrino fluxes
(b) possible need of significant hot component \cite{dm} in the dark
matter of the universe and
(c) some indication of $\bar{\nu}_e-\bar{\nu}_{\mu}$ oscillations in
the laboratory \cite{lsnd}.
These hints can be reconciled with each
other if there exists a fourth very light ($< {\cal O}$(eV))
neutrino mixed with some of the known neutrinos preferably
with the electron one. The fourth neutrino is
required to be sterile in view of the strong bounds on number of
neutrino flavours coming both from the LEP experiment
and from the primordial nucleosynthesis \cite{ns}.
The existence of a very light sterile neutrino demands theoretical
justification since unlike the active neutrinos, the mass of a
sterile state is not protected by the gauge symmetry of the SM
and hence could be very large. Usually a sterile neutrino is
considered on the same footing as the active neutrinos and some
ad hoc symmetry is introduced to keep this neutrino light.
Recently there are several attempts to construct models for sterile
neutrinos which have their origin beyond the usual lepton
structure \cite{paper1,paper2,mirror1,mirror2,ma}.
In this report, we discuss the role of supersymmetry (SUSY) in explaining
both the existence and the lightness of a singlet fermion $S$ which can mix
with the neutrinos. As a case of special interest we will concentrate
on the mass of $S$ and its mixing with the electron neutrino in the range:
\begin{eqnarray} \label{parameters}
m_S &\simeq& (2-3)\cdot 10^{-3} \,{\rm eV} \nonumber\\
\sin\theta_{es} &\simeq& \tan\theta_{es} \simeq (2-6)\cdot 10^{-2} \;.
\end{eqnarray}
These values of parameters allow one to solve the solar neutrino
problem through the resonance conversion $\nu_e \to S$ \cite{msw}.
More discussions on simultaneous reconciliations of the diverse neutrino
problems can be found in refs.~\cite{paper1,paper2} on which this report is
based.
\section{Quasi Goldstone Fermion}
The existence of SM-singlet fields is a common property in physics
beyond the standard model.
The most interesting examples are the Goldstone bosons of spontaneously
broken global symmetries required to solve the strong CP problem (the
Peccei-Quinn symmetry) \cite{pq} and to explain the origin of neutrino masses
(the lepton number symmetry) \cite{cmp}.
In the SUSY limit, a spontaneously broken global
symmetry automatically generates a massless singlet (Goldstone) fermion
being a superpartner of a Goldstone boson. However, SUSY breakdown
results in generation of mass of a Goldstone fermion.
While the existence of these quasi Goldstone fermions (QGF)
is logically independent of neutrino physics, there are good
reasons to expect that these fermions will couple to neutrinos.
Indeed, in the case of lepton number symmetry the superfield which is
mainly responsible for the breakdown of the lepton number symmetry
carries nontrivial lepton number and therefore it can directly couple
to leptons if the charge is appropriate.
In the case of the PQ symmetry,
this superfield could couple to the Higgs supermultiplet.
If theory contains small violation of $R$-parity then this mixing
with the Higgs gets communicated to the neutrino sector.
Thus the occurrence of a QGF can have implications for neutrino
physics.
In the following subsections we elaborate upon the expected properties
of the QGF: their masses arising after SUSY breaking and
the mixing of these fermions with the electron neutrino.
\subsection{masses of QGF}
The supersymmetric standard model with some global symmetry $U(1)_G$ can be
characterized by the following superpotential:
\begin{equation} \label{w}
W=W_{MSSM}+W_S+W_{mixing} \;, \end{equation}
where $W$ is assumed to be invariant under $U(1)_G$.
As we outlined in the above, this symmetry may be identified with the
PQ symmetry, lepton number symmetry or combination thereof. The
first term in eq.~\refs{w} refers to the superpotential of the minimal
supersymmetric standard model (MSSM).
The second term contains $SU(3)\times SU(2)\times U(1)$ singlet superfields which are responsible
for the breakdown of $U(1)_G$.
The minimal choice for $W_S$ is
\begin{equation} \label{ws} W_S=\lambda (\sigma \sigma'- f_G^2) y \;, \end{equation}
where $\sigma,\sigma'$ carry non trivial $G$-charges and $f_G$ sets the
scale of $U(1)_G$ breaking.
The last term of eq.~\refs{w} describes mixing of the singlet
fields with the superfields of the MSSM.
In the case \refs{ws} the Goldstone fermion is contained in
$S\sim \sigma-\sigma'$ and is massless in the SUSY limit.
Broken SUSY itself cannot automatically protect the mass of a QGF.
It depends on the structure of the superpotential
$W_S$~\cite{chun1} and on the pattern of soft-terms~\cite{chun2}.
It also depends on the way this breaking is communicated to the singlet $S$
and the scale $f_G$ \cite{paper2}.
The most natural framework for light QGF is no-scale
supergravity~\cite{noscale}. No-scale models
contain only one kind of soft-terms, namely, gaugino masses.
Therefore, th soft SUSY-breaking terms corresponding to $W_S$ in eq.~\refs{w}
are absent at tree-level and thus QGF remains massless.
However, the radiative mass can be triggered by the $SU(3)\times SU(2)\times U(1)$ gaugino masses
through a set of interactions.
A realistic example can be found in the context of the seesaw mechanism.
The vacuum expectation value (VEV) of the field
$\sigma$ (or $\sigma'$) may give rise to large masses of right-handed (RH)
neutrinos $N$ as in the following superpotential invariant under
$U(1)_G$:
\begin{equation}\label{seesaw}
W = {m^D \over \langle H_2 \rangle} L N H_2 + \frac{M}{f_{G}} N N \sigma \;, \end{equation}
where we have omitted the generation indices.
The generation structure of the superpotential \refs{seesaw} will depend on
the $U(1)_G$-charge assignment to the fields \cite{paper2}.
This $U(1)_G$ symmetry is not necessarily the lepton number symmetry as we
will discuss in subsection 2.2.
The first term in eq.~\refs{seesaw} gives rise to the Dirac masses
of the neutrinos, whereas the second one gives the Majorana masses of
RH neutrino components.
The scale $f_{G} \sim 10^{10} - 10^{12} \,{\rm GeV}$ generates
$M \sim 10^{10}-10^{11}$ GeV required by the hot dark matter
and atmospheric neutrinos.
If the soft-term $A_N NN\sigma$ with $A_N \sim m_{3/2}$ is present, there appears
one-loop mass of the QGF proportional to $A_N$ \cite{ckl}.
But in no-scale models $A_N=0$ at tree-level and the QGF mass is
indeed generated in three loops as shown in Figure~1.
\begin{figure}
\begin{picture}(200,150)(-220,-20)
\put(-80,0){\line(160,0){160}}
\multiput(0,0)(0,-2.5){10}{\circle*{.1}}
\multiput(-50,0)(0,2.5){30}{\circle*{.1}}
\multiput(50,0)(0,2.5){30}{\circle*{.1}}
\put(-50,75){\line(4,3){50}}
\put(-50,75){\line(4,-3){50}}
\multiput(50,75)(-2,1.5){25}{\circle*{.1}}
\multiput(50,75)(-2,-1.5){25}{\circle*{.1}}
\multiput(50,75)(2,1.5){10}{\circle*{.1}}
\put(0,37.5){\line(0,75){75}}
\put(0,75){\makebox(0,0){$\times$}} \put(3,73){$m_{1/2}$}
\put(70,90){\makebox(0,0){$\times$}} \put(67.5,81){$\sigma$}
\put(0,-25){\makebox(0,0){$\times$}} \put(3,-27){$\sigma$}
\put(-80,5){\makebox(0,0){$S$}}
\put(80,5){\makebox(0,0){$S$}}
\put(-25,5){\makebox(0,0){$N$}}
\put(25,5){\makebox(0,0){$N$}}
\put(-28,99){\makebox(0,0){$L$}}
\put(28,99){\makebox(0,0){$L$}}
\put(-28,50){\makebox(0,0){$H_2$}}
\put(29,49){\makebox(0,0){$H_2$}}
\put(-62,37.5){$N$}
\put(52,37.5){$N$}
\end{picture}
\caption{Three-loop diagram for the QGF mass.
The cross with $m_{1/2}$ denotes gaugino mass insertion.}
\end{figure}
This three-loop mass can be estimated as
\begin{equation} \label{radmass3}
m_S \simeq {\alpha_2\over (4\pi)^5} {m_{\nu} M^3 \over v_2^2 f_{G}^2} m_{1/2}
\;. \end{equation}
Here $\alpha_2$ and $m_{1/2}$ are the $SU(2)$ fine structure constant and
gaugino mass respectively.
For $m_\nu \simeq 3$ eV, $m_{1/2} \simeq v_2 \simeq 100$ GeV, and $f_G \simeq
10^{12}$ GeV, one gets $m_S \simeq 3 \cdot 10^{-3}$ eV with a value of
$M \simeq 10^{10}$ GeV.
A contribution to the mass of the QGF can follow also from
interactions, $W_{mixing}$, which mix $S$ with usual neutrinos
(subsection 2.2).
\subsection{Neutrino-QGF mixing}
We now discuss how the QGF can mix with neutrinos.
Such a mixing implies the violation of $R$-parity
conventionally imposed in the MSSM \cite{hall}. This is simply because that
the leptons being ordinary matter fields are $R$-even and the QGF being a
fermionic partner of a Goldstone boson is $R$-odd.
The violation of $R$-parity may destabilize the lightest supersymmetric
particle (LSP) which is usually considered as the cold dark matter (CDM)
of the Universe.
For this reason, we consider the PQ symmetry as a good candidate for
$U(1)_G$ since the coherent oscillation of the axion can provide the CDM
for $f_{PQ} \sim 10^{12}$ GeV \cite{kt}.
Therefore, the PQ mechanism required for a resolution of the strong CP
problem can supply both the CDM and the sterile neutrino.
The best way to implement the PQ symmetry in the MSSM is to extend the Higgs
mass term in such a way that the smallness of the Higgs mass parameter $\mu$
can be naturally obtained. For instance, let us consider the
non-renormalizable term \cite{mu}
\begin{equation} \label{nr}
\lambda H_1H_2\frac{\sigma^2}{M_P}\;, \end{equation}
where $M_P$ is the Planck mass \footnote{One can also introduce the
renormalizable term to generate $\mu \simeq m_{3/2}$ \cite{chun3}.}.
Here the VEV of $\sigma$, $\langle \sigma \rangle
\sim f_{PQ}$, spontaneously breaks the PQ symmetry.
In this case, $\mu= \lambda\frac{\langle\sigma\rangle^2}{M_P}$ can be about the weak
scale. When the axion superfield $S$ is predominantly consists of $\sigma$,
the PQ symmetry breaking yields the Higgs mass term and the coupling of $S$
to the Higgs superfields
\begin{equation}\label{mix1} W_{mixing}=c_\mu\frac{\mu}{f_{PQ}}H_1H_2S + \mu H_1H_2 \end{equation}
with $c_\mu$ being ${\cal O}(1)$.
In order to have the mixing of $S$ with neutrinos, one needs the lepton
number violating term $\epsilon LH_2$.
It is remarkable to notice that the PQ scale is in the right range for
the RH neutrino masses.
The PQ symmetry can indeed play a role of the lepton number symmetry if
both the Higgs and leptons transform non-trivially under the PQ symmetry
as in ref.~\cite{lpy}. In this case one can correlate
the origin of $\epsilon$ and $\mu$ to the same symmetry breaking scale $f_{PQ}$.
The neutrino and Higgs coupling to QGF is then given by
\begin{eqnarray}\label{mix3}
W_{mixing}&=&\mu H_1H_2+\epsilon L_eH_2 + \nonumber\\
& & c_{\mu}\frac{\mu}{f_{PQ}}H_1H_2S +
c_{\epsilon}\frac{\epsilon}{f_{PQ}}L_eH_2S \;, \end{eqnarray}
where $L_e$ is the electron doublet.
If the PQ symmetry is the standard one unrelated to
the lepton sector, the parameter $\epsilon$ vanishes.
On the other hand, the global $U(1)$ symmetry becomes
the usual lepton number symmetry when $c_\mu = 0$ and the bare $\mu$-term is
introduced.
An example of models which leads to the mixing terms of eq.~\refs{mix3}
can be obtained by the PQ-charge prescription ($-1$,$-1$, 1,$-1$,$-2$) for
($H_1$, $H_2$, $\sigma$, $\sigma'$, $L_e$). It permits
the following $U(1)_{PQ}$ invariant superpotential:
\begin{equation} \label{model2}
W = \lambda (\sigma\s' - f_{PQ}^2)y + {\delta_\mu \over M_P} H_1 H_2 \sigma^2
+{\delta_\epsilon \over M_P^2} L_e H_2 \sigma^3 \;, \end{equation}
which gives the terms displayed in eq.~(\ref{mix3}) with
$c_\epsilon={3\over\sqrt{2}},c_\mu=\sqrt{2}$.
\smallskip
The $W_{mixing}$ in eq.~\refs{mix3}
generates the following effective mass matrix for $\nu_e$ and $S$
\begin{equation}\label{matrix3}
\left( \begin{array}{cc}
0&(c_\epsilon-c_\mu) \epsilon v\sin\beta/ f_{PQ}\\
(c_\epsilon-c_\mu) \epsilon v\sin\beta/ f_{PQ}
&m_S^0- c_{\mu}^2\mu v^2 \sin2\beta / f_{PQ}^2\\
\end{array} \right) \;, \end{equation}
where we added the direct mass $m_S^0$ which can be generated by the
mechanism of subsection 2.1.
According to eq.~\refs{matrix3} the $\nu_e-S$ mixing angle $\theta_{es}$
is determined by
\begin{equation}\label{ts2}
\tan \theta_{es}\sim \frac{(c_\mu-c_\epsilon) \epsilon v\sin\beta}
{m_S^0 f_{PQ}- c_{\mu}^2\mu v^2 \sin2\beta / f_{PQ}} \;. \end{equation}
For $f_{PQ} \simeq 10^{12}$ GeV, $m_S^0 \simeq 3\cdot 10^{-3}$ eV is the
dominant contribution to the mass of $S$. In this case one obtains from
eq.~(\ref{ts2}) for the $\nu_e-S$ mixing
\begin{equation} \label{ts1}
\tan \theta_{es}\sim \frac{\epsilon v \sin \beta}
{m_S^0 f_{PQ}}\;. \end{equation}
Then the desired value, $\tan \theta_{es} \sim (2 - 6)\cdot 10^{-2} \,{\rm eV}$
\refs{parameters}, can be obtained if the $R$-parity breaking parameter
$\epsilon$ equals
\begin{equation} \label {epsis}
\epsilon \sim \frac{m_S^0 f_{PQ} \tan \theta_{es}}{v \sin\beta}
\approx (2 - 6)\cdot 10^{-16} \frac{f_{PQ}}{\sin \beta} \;. \end{equation}
For $f_{PQ} \sim 10^{12}$ GeV one has $\epsilon \sim 0.1$ MeV.
Let us remark the other possibilities for the QGF mass.
If $m_S^0 = 0$ in eq.~\refs{matrix3}, the QGF mass,
$m_S = (2 - 3)\cdot 10^{-3} \,{\rm eV}$
can be obtained for the marginally allowed value of the PQ scale:
\begin{equation} f_{PQ} \approx v \sqrt{\frac{\mu \sin 2\beta}{m_S}}
\stackrel{\scriptstyle <}{\scriptstyle\sim} 4 \cdot10^9 \,{\rm GeV} \;. \end{equation}
For $f_{PQ} > 10^{10}$ GeV the QGF mass generated via $\mu$-term
is too small for the MSW solution. For $f_{PQ} \sim 10^{11}$ GeV,
$m_S \approx 10^{-5} \,{\rm eV}$ is in the region of ``just-so" solution
of the solar neutrino problem.
In these cases, however, axions cannot provide the CDM as we noted before.
\section{A light singlet in the standard seesaw structure}
In the previous case, the QGF mixes with the electron neutrino directly ($\epsilon
c_\epsilon \neq 0$) or via its coupling to the Higgses ($c_\mu \neq 0$). The small
mass of the QGF was related to the multi-loop effect or the suppression by
$1/f_{PQ}^2$ due to the Goldstone property. An important consequence was the
$R$-parity violation leading to destabilization of the LSP.
In this section, we will suggest another scheme in which $R$-parity is
preserved. For this, one should place the singlet $S$ in the superfield with
zero VEV. This implies that the singlet has to be introduced from outside.
Being a singlet $S$ can mix with neutrinos via its coupling to the
right-handed neutrinos. In this case, the existence of $S$ cannot be
explained but the smallness of its mass can be understood in terms of
the seesaw mechanism. In order to implement a light singlet fermion in the
standard seesaw structure, we will suggest to use $R$-symmetry which
occurs in many SUSY theories. The (unbroken) $R$-parity is then
embedded in the $R$-symmetry.
\smallskip
Let us first determine the parameters appearing in the phenomenological
superpotential
\begin{equation} \label{base}
W = {m_e \over \langle H_2 \rangle}L_e N_eH_2 + {M_e\over 2}N_eN_e
+ m_{es}N_e S \;, \end{equation}
where $N_e$ is the right-handed neutrino component.
The Dirac mass $m_e$ and the mixing mass $m_{es}$
are much smaller than the Majorana mass $M_e$: $m_e, m_{es} \ll M_e$.
The superpotential \refs{base} leads to the mass matrix in the basis
$(S, \nu_e, N_e)$:
\begin{equation} \label{mm1}
{\cal M} = \left(\ba{ccc} 0 & 0 & m_{es}\\ 0 & 0 & m_e \\ m_{es} & m_e & M_e
\ea\right) \;.
\end{equation}
The diagonalization of \refs{mm1} is straightforward. One combination of
$\nu_e$ and $S$ is massless and the orthogonal combination
acquires a mass via the see-saw mechanism:
\begin{equation} \label{m1}
m_1 \simeq -{m_e^2 + m_{es}^2 \over M_e}\;.
\end{equation}
The mass of the heavy neutrino is $\simeq M_e$.
The $\nu_e$--$S$ mixing angle is determined by
\begin{equation} \label{th}
\tan\theta_{es} = {m_e \over m_{es}} \;.
\end{equation}
Taking for $m_e$ the typical Dirac mass of the first generation:
$m_e \sim (1-5) \,{\rm MeV}$, and suggesting that $\nu_e \rightarrow
S$ conversion explains the solar neutrino problem with $m_1 =m_S$
as in \refs{parameters}, we find
\begin{equation} \label{mis}
m_{es} = {m_e \over \tan\theta_{es}} \simeq (0.02-0.3) \,{\rm GeV} \;.
\end{equation}
According to \refs{m1} the RH mass scale is
\begin{equation} \label{Mis}
M_e \simeq m_{es}^2/m_1 = {m^2_e \over m_1 \tan^2\theta_{es}}
\simeq (10^8-3\cdot10^{10})\,{\rm GeV}\;.
\end{equation}
One has now to understand how the mixing mass \refs{mis} arises without
introducing new mass scales.
One also has to ensure that there is no direct coupling of $S$ with $L_e$,
and the mass term $SS$ is absent or negligibly small.
\smallskip
Our prescription is quite simple. Consider the superpotential
\begin{equation} \label{model3}
W = {m_e\over \langle H_2\rangle} L_e N_eH_2 + f N_e N_e\sigma + f' N_e S y
- {\lambda \over 2}(\sigma^2 - M^2)y \;. \end{equation}
whose structure is determined by the $R$-symmetry
under which the fields ($L_e$, $N_e$, $S$, $y$, $\sigma$, $H_2$) carry
the $R$-charges (1, 1,$-1$, 2, 0, 0).
Note that the $R$-symmetry forbids the bare mass terms $SS$ as well as
the coupling $SS\sigma$. The last term in eq.~\refs{model3} can be replaced by
$(\sigma\s'-M^2)y$ to implement the lepton number symmetry.
In the global SUSY limit, $\sigma$ gets non-zero
VEV $\langle \sigma \rangle \simeq M \sim 10^{11}$ GeV which generates the Majorana
mass of $N_e$: $M_{e} = f \langle \sigma \rangle$.
The point is that $y$ develops a VEV as a consequence of SUSY breaking.
Broken SUSY produces the following soft-breaking terms in the scalar
potential:
\begin{eqnarray} \label{soft}
V_{soft} &=& \{ A_L {m_e \over \langle H_2 \rangle} L_e N_eH_2 +
fA_\nu N_e N_e\sigma + f' A_S N_eSy - \nonumber\\
& & {\lambda \over2}(A_y\sigma^2 - B_yM^2)y + \mbox{h.c.} \} +
\sum_i m_i^2 |z_i|^2 \;, \end{eqnarray}
where $z_i$ denotes the fields appearing in the superpotential \refs{model3}
and $A_L$, etc., are the soft-breaking parameters.
Minimization of the potential shows the following:
(1) The fields $L_e, N_e, S$ do not develop VEV and therefore $R$-parity is
unbroken.
(2) The field $y$ acquires non-zero VEV due to the soft-breaking terms.
Consequently, the mixing mass for $S$ and $N_e$ appears:
\begin{equation} \label{mis2}
m_{es}= {f'\over 2\lambda}(A_y-B_y) \end{equation}
Since $m_{es} \gg m_1$, no strong tunning of $A_y-B_y$ is needed.
For $A_y-B_y \sim {\cal O}(m_{3/2})$,
the desired value of $m_{es}$ \refs{mis}
can be obtained by choosing $f'/\lambda \sim 10^{-3}-10^{-2}$.
However, more elegant possibility is that $A_y=B_y$ at tree level
but a non-zero value for $A_y - B_y$ is generated due to radiative
corrections through the differences in interactions of $\sigma$ and $y$.
In this case one expects
\begin{equation} \label{mrad}
m_{es} \sim {\bar{\lambda}^2\over 16\pi^2} m_{3/2} \;,
\end{equation}
where $\bar{\lambda}$ represents a combination of the constants $\lambda,f$ and $f'$.
As a consequence, the value $m_{es}\sim 0.1$ GeV does not require
smallness of $\bar{\lambda}$ or $f'$.
The equality $A_y = B_y$ at tree level can be achieved by the
introduction of non-minimal K\"ahler potential allowing
mixings between the observable and hidden sectors.
Let us introduce the following K\"ahler potential:
\begin{equation}
K = C\overline{C} + C\overline{C}(a\frac{Z}{M_{Pl}} +
\overline{a}\frac{\overline{Z}}{M_{Pl}}) + Z\overline{Z} \;,
\end{equation}
where $C$ and $Z$ represent an observable and hidden sector field,
respectively. Then usual assumption that the observable sector has no
direct coupling to the hidden sector in superpotential, $W=W(C) + W(Z)$,
leads to the universal soft-terms:
\begin{equation}
V_{soft} \sim m_{3/2} W(C) + \mbox{h.c.} \;,
\end{equation}
provided $\overline{a}= \langle W(Z) \rangle / \langle M_{Pl} \partial W/\partial Z +
W(Z)\overline{Z}/M_{Pl} \rangle $.
Note also that the field $C$ does not acquire a soft-breaking mass.
This mechanism can be generalized to arbitrary number of observable sector
fileds. For our purpose $C \equiv \sigma, y$, i.e., we couple $\sigma$ and $y$ to
the hidden sector field $Z$ with the above-mentioned choice for $a$.
\section{Conclusions}
Simultaneous presence of different neutrino anomalies points to the
existence of a sterile neutrino. In particular, the resonance conversion
of the electron neutrino into such a singlet fermion $S$ can explain
the solar neutrino problem provided its mass and mixing are appropriate
\refs{parameters}.
Supersymmetry is shown to provide a framework within which the existence
and the desired properties of such a light fermion follow naturally.
We have considered first a possibility that the sterile neutrino is
a quasi Goldstone fermion appearing in supersymmetric theories
as a result of spontaneous breaking of a global $U(1)_G$ symmetry.
This global $U(1)_G$ symmetry can be identified with the PQ
symmetry, the lepton number symmetry.
The smallness of $m_S$ can be attributed in supergravity theory
to no-scale kinetic terms for certain superfields.
The mixing of QGF with the neutrinos implies spontaneous or explicit
violation of $R$-parity. QGF can mix with neutrino via interaction with
Higgs multiplets (in the case of PQ symmetry) or directly via coupling with
the combination $L H_2$ (in the case of lepton number symmetry).
In the case of the PQ symmetry, the PQ-scale $f_{PQ}\sim 10^{10}-10^{12}$ GeV
determines several features of the model presented here.
It provides simultaneous explanation of the parameters $\epsilon$ and $\mu$ and
thus leads to small $R$-parity violation ($\epsilon LH_2$ with $\epsilon \sim 0.1$ MeV)
required in order to solve the solar neutrino problem in our approach.
It also provides the intermediate scale
for the right-handed neutrino masses which is required in order to solve
the dark matter and the atmospheric neutrino problem. Furthermore,
it controls the magnitude of the radiatively generated mass of the QGF
and allows it to be in the range needed for the
MSW solution of the solar neutrino problem.
Finally, the CDM can consist of the axion if $f_{PQ} \sim 10^{12}$ GeV.
Thus the basic scenario presented here is able to correlate variety of
phenomena.
The conservation of R-parity requires for the fermion $S$ to be
a component of singlet superfield which has no VEV.
This allows to construct simple model \refs{model3} in which the
properties (mass and mixing) of $S$ follow from the conservation of
$R$-symmetry. The singlet field is mixed with RH neutrinos by the
interaction with the field $y$ which can acquire VEV radiatively after soft
SUSY breaking.
Let us finally comment on the other phenomenological consequences of the
existence of such a sterile state $S$.
An $U(1)_G$ symmetry being generation-dependent \cite{paper1,paper2}
can provide simultaneous explanations for the predominant coupling of $S$ to
the first generation (thus satisfying the nucleosynthesis bound)
and for the pseudo-Dirac structure of $\nu_\mu$--$\nu_\tau$ needed in solving
the atmospheric neutrino and the hot dark matter problem.
In this case, it appears nontrivial to accommodate the
parameters of $\bar{\nu}_\mu \to \bar{\nu}_e$ oscillations in the region of
sensitivity of LSND and KARMEN experiments.
The simplest way is to introduce a slight violation of the $U(1)_G$ symmetry
through which such parameters can be incorporated.
|
\section{Quantum Calogero-Sutherland model}\label{QCSM}
Define $N$ differential operators $\{H_k\}_{k=1}^N$,
acting on functions of $N$ variables $\vec q=\{q_1,\ldots,q_N\}$
and depending on a parameter $g$, by the formula \cite{Oshima-Seki}
\begin{equation}\label{eq:def-Hk}
H_k=\sum_{0\leq l\leq\frac{k}{2}}\sum_{\sigma\in\mbox{\frak S}_N}
\frac{1}{\# G(l,k-2l)}D^\sigma_{l,k-2l}
\end{equation}
where
\begin{equation}
D_{m,n}=u(q_1-q_2)u(q_3-q_4)\ldots u(q_{2m-1}-q_{2m})
\frac{(-i)^n\partial^n}{\partial q_{2m+1}\partial q_{2m+2}\ldots\partial q_{2m+n}}.
\end{equation}
Here we denote $u(q)=-g(g-1)/\sin^2 q$, whereas $\mbox{\frak S}_N$ is the
permutation
group of the set $\{1,\ldots,N\}$, and $G(m,n)=\{\sigma\in\mbox{\frak S}_N |
D^\sigma_{m,n}=D_{m,n}\}$.
Note that, when $g\rightarrow0$, the operators $H_k$ behave as
\begin{equation}
H_k=(-i)^k\sum_{j_1<\cdots<j_k}\frac{\partial^k}{\partial q_{j_1}\ldots\partial q_{j_k}}
+{\cal O}(g),
\end{equation}
providing thus a one-parameter deformation of the elementary symmetric
polynomials in $\partial/\partial q_j$.
It is known \cite{Oshima-Seki} that the operators $H_k$ generate
a commutative ring which contains, in particular, the quantum
Calogero-Sutherland \cite{Calog,Suth,OP1,OP2} Hamiltonian
\begin{equation}
H=\frac{1}{2} H_1^2-H_2=-\frac{1}{2}\sum_{j=1}^N\frac{\partial^2}{\partial q_j^2}
+\sum_{j_1<j_2}\frac{g(g-1)}{\sin^2(q_{j_1}-q_{j_2})}.
\end{equation}
To describe the quantum problem more
precisely, define the space of quantum states ${\cal H}^{(N)}$ as the
complex Hilbert space of functions $\Psi$ on the torus
$T^{(N)}=\hbox{\bbd R}} %% \def\R{{\bf R}^N/\pi\hbox{\bbd Z}^N\ni\vec q$ which are symmetric w.r.t.\ the permutations
of $q_j$, the scalar product being defined as
\begin{equation}\label{eq:norm}
\left<\Psi,\Phi\right>=\int_0^\pi \!dq_1\ldots\int_0^\pi \!dq_N
\,\bar\Psi(\vec q)\Phi(\vec q).
\end{equation}
Note that for the real $g$
the operators (\ref{eq:def-Hk}) are formally Hermitian
w.r.t.\ the above sesquilinear form.
Let the vacuum (ground state) function $\Omega$ be defined as
\begin{equation}\label{eq:Omega}
\Omega(\vec q)=\left|\prod_{j<k}\sin(q_j-q_k)\right|^g.
\end{equation}
Though $\Omega\in{\cal H}^{(N)}$ for $g>-\frac{1}{2}$, we shall assume a more
strong condition $g>0$ which simplifies description of the eigenvectors.
Let ${\cal T}^{(N)}$ be the space
of symmetric trigonometric polynomials in variables $\vec q$,
that is the symmetric Laurent polynomials in variables
$t_j=e^{2iq_j}$. The simplest way to fix the ``boundary
conditions'' for the operators $H_k$ is to restrict them first on the
dense linear subset ${\cal D}_g^{(N)}=\Omega{\cal T}^{(N)}\subset{\cal H}^{(N)}$.
Since ${\cal D}_g^{(N)}$ consists
of common analytical vectors of operators $H_k$, the latter can be
extended uniquely to commuting self-adjoint operators in ${\cal H}^{(N)}$.
The complete set of orthogonal eigenvectors to the self-adjoint $H_k$
\begin{equation}
H_k\Psi_{\vec n}=h_k\Psi_{\vec n}
\end{equation}
is well known \cite{Suth,OP2}.
The eigenvectors are parametrized by the sequences
$\vec n=\{n_1\leq n_2\leq\ldots \leq n_N\}$
of integers $n_j\in\hbox{\bbd Z}$. The corresponding eigenvalues $h_k$ are
\begin{equation}\label{eq:def-hm}
h_k=2^k\sum_{j_1<\cdots<j_k}m_{j_1}\ldots m_{j_k},
\qquad m_j=n_j+g\left(j-\frac{N+1}{2}\right).
\end{equation}
The eigenfunctions allow the factorization
\begin{equation}\label{eq:Psi-Om}
\Psi_{\vec n}({\vec q})=\Omega(\vec q)J_{\vec n}({\vec q}),
\qquad J_{\vec n}\in{\cal T}^{(N)}.
\end{equation}
In particular, for the ground state $\Omega=\Psi_{0\ldots0}$
and $J_{0\ldots0}=1$.
The symmetric trigonometric polynomials $J_{\vec n}$ are known
as Jack polynomials corresponding to the root system $A_{N-1}$
or simple Lie algebra $sl_N$, see \cite{Macd} and also \cite{Koor74} for
the $A_2$ case.
Our notation differs slightly from the conventional one: our
parameter $g$ relates to $\alpha$ used in \cite{Macd}
as $g=\alpha^{-1}$,
and we do not impose the restriction $n_j\geq 0$.
The problem of finding square integrable eigenfunctions
$\Psi\in{\cal H}^{(N)}$
of the operators $H_k$ turns out thus to be equivalent to
the purely algebraic problem of finding the polynomial eigenfunctions
$J\in{\cal T}^{(N)}$ of the differential operators $\tilde H_k$
obtained by conjugation of $H_k$ with $\Omega$
\begin{equation}\label{eq:def-tHk}
\tilde H_k=\Omega^{-1}H_k\Omega.
\end{equation}
Jack polynomials can be considered as a one-parametric deformation
of elementary symmetric polynomials
$ S_{\vec n}(\vec q)=\sum t_1^{\nu_1}\ldots t_N^{\nu_N} $
where the sum is taken over all distinct permutations $\vec \nu$
of $\vec n$, such that
\begin{equation}\label{eq:norm-J}
J_{\vec n}=S_{\vec n}+\sum_{\vec n'\preceq \vec n}
\kappa_{\vec n\vec n'}S_{\vec n'},
\end{equation}
where $\kappa_{\vec n\vec n'}$ is a rational function in $g$ vanishing
for $g=0$, and the dominant order for sequences $\vec n$ is defined as
\begin{equation}\label{eq:dominant}
\vec n\succeq\vec n' \quad\Longleftrightarrow\quad
\left\{\sum_{j=1}^N n_j=\sum_{j=1}^N n_j^\prime; \quad
\sum_{j=k}^N n_j \geq \sum_{j=k}^N n_j^\prime, \quad
k=2,\ldots,N \right\}
\end{equation}
Another important property of Jack polynomials is the orthogonality
with the weight $\Omega^2$,
\begin{equation}
\int_0^\pi \!dq_1\ldots\int_0^\pi \!dq_N \,
\bar J_{\vec n}(\vec q)J_{\vec n'}(\vec q)\Omega^2(\vec q)=0,
\quad \vec n\neq \vec n'
\end{equation}
For the generalization of Jack polynomials for other root systems see
\cite{HeckOp}.
\section{Separation of variables: conjectures}\label{SoV}
In the classical case, when the differentiation $-i\partial/\partial q_j$ is
replaced by the momentum $p_j$ canonically conjugated to $q_j$,
the Calogero-Sutherland system is completely integrable in the
Liouville's sense \cite{Calog,OP1}. It is thus natural
to speak of its quantum version described above as a quantum integrable
system. The common property to be expected from an integrable system
(classical or quantum one) is the {\it separability of variables}
\cite{Kaln,Kuz,Skl:32,Skl:38} which suggests the following
conjecture.
{\bf Conjecture 1.} {\it There exists a linear operator
\begin{equation}
K:\Psi_{\vec n}({\vec q})\longmapsto
\tilde\Psi_{\vec n}(y_1,\ldots,y_{N-1};Q)
\end{equation}
such that any eigenfunction $\Psi_{\vec n}$ is transformed into
the factorized function
\begin{equation}
\label{eq:factor-Psi}
\tilde\Psi_{\vec n}(y_1,\ldots,y_{N-1};Q)=
e^{ih_1Q}\prod_{k=1}^{N-1}\psi_{\vec n}(y_k).
\end{equation}
The distinguished variable $Q\equiv q_N$ is simply the coordinate canonically
conjugated to the total momentum $H_1$.
The study of the low-dimensional cases $N=2,3$
allows to formulate an even more detailed conjecture about
the structure of the separated eigenfunction $\tilde\Psi$.
{\bf Conjecture 2.} {\it The factor $\psi_{\vec n}(y)$ in
(\ref{eq:factor-Psi}) allows further factorization
\begin{equation}\label{eq:fact-psi}
\psi_{\vec n}(y)=(\sin y)^{(N-1)g}\varphi_{\vec n}(y)
\end{equation}
where $\varphi_{\vec n}(y)$ is a Laurent polynomial in $t=e^{2iy}$
\begin{equation}\label{eq:phi-Laurent}
\varphi_{\vec n}(y)=\sum_{k=n_1}^{n_N}t^k c_k(\vec n;g).
\end{equation}
The coefficients $c_k(\vec n;g)$ are rational functions of
$k$, $n_j$ and $g$. Moreover, $\varphi_{\vec n}(y)$
can be expressed
explicitely in terms of the hypergeometric function ${}_N F_{N-1}$
as
\begin{equation}\label{eq:phi-hgf}
\varphi_{\vec n}(y)=t^{n_1}(1-t)^{1-Ng}
{}_N F_{N-1}(a_1,\ldots,a_N;b_1,\ldots,b_{N-1};t)
\end{equation}
where
\begin{equation}\label{eq:def-ab}
a_j=n_1-n_{N-j+1}+1-(N-j+1)g, \qquad
b_j=a_j+g,
\end{equation}
\begin{equation}\label{eq:def-hgf}
{}_N F_{N-1}(a_1,\ldots,a_N;b_1,\ldots,b_{N-1};t)=
\sum_{k=0}^\infty\frac{(a_1)_k\ldots(a_N)_k t^k}%
{(b_1)_k\ldots(b_{N-1})_k k!},
\end{equation}
and $(a)_k$ is the standard Pochhammer symbol:
\begin{equation}
(a)_0=1, \qquad (a)_k=a(a+1)\ldots(a+k-1)=
\frac{\Gamma(a+k)}{\Gamma(a)}.
\end{equation}
The conjectures 1 and 2 are proved
in the next section for the $N=2$ case
and in the sections \ref{int-tr} and \ref{sep-eq} for the $N=3$ case.
Section \ref{sep-eq} contains also
a more detailed discussion of the conjecture 2 for $N>3$, see theorem
\ref{Vadim}.
Further support to the conjectures is given by the study
of the case $g=1$ when Jack polynomials degenerate into Schur functions
(section \ref{Schur}).
\section{$A_1$ case}
It is a well known fact that in the $A_1$ case Jack polynomials
are reduced to hypergeometric polynomials of one variable
\cite{HeckOp}.
Nevertheless, we review the derivation briefly in order to
prepare the stage for the discussion of the $A_2$ case.
For $N=2$ the commuting operators (\ref{eq:def-Hk}) are
\begin{equation}
H_1=-i(\partial_1+\partial_2), \qquad
H_2=-\partial_1\partial_2-g(g-1)\sin^{-2}q_{12}.
\end{equation}
(we denote $\partial_j=\partial/\partial q_j$ and $q_{jk}=q_j-q_k$).
Respectively,
\begin{eqnarray*}
\tilde H_1=-i(\partial_1+\partial_2), \qquad
\tilde H_2=-\partial_1\partial_2+g\cot q_{12}(\partial_1-\partial_2)-g^2,
\end{eqnarray*}
the vacuum vector being
\begin{equation}
\Omega(\vec q)=\left|\sin q_{12}\right|^g.
\end{equation}
The eigenvectors $\Psi_{\vec n}$, resp.\ $J_{\vec n}$,
according to (\ref{eq:def-hm}), are parametrized
by the pairs of integers $\vec n=\{n_1\leq n_2\}$, the corresponding
eigenvalues being
\begin{equation}
h_1=2(m_1+m_2)=2(n_1+n_2), \qquad
h_2=4m_1m_2=(2n_1-g)(2n_2+g)
\end{equation}
where
\begin{equation}
m_1=n_1-\frac{g}{2}, \qquad m_2=n_2+\frac{g}{2}.
\end{equation}
The separation of variables is given by the simple change of coordinates
\begin{equation}
K:\Psi(q_1,q_2) \longmapsto \tilde\Psi(y,Q)=\Psi(y+Q,Q).
\end{equation}
Actually, the calculations would be simpler for the more
symmetric definition $Q=(q_1+q_2)/2$ rather than $Q=q_2$ but
we wish to preserve here the coherence of notation for
the study of $N=3$ case.
The spectral problem $H_k\Psi=h_k\Psi$ rewritten in terms
of the function $\tilde\Psi$ reads
\begin{equation}
\left[\partial_Q-ih_1\right]\tilde\Psi=0, \qquad
\left[\partial^2_y-\partial_y\partial_Q-\frac{g(g-1)}{\sin^2 y}-h_2\right]\tilde\Psi=0,
\end{equation}
allowing immediate separation of variables of the form (\ref{eq:factor-Psi})
\begin{equation}
\tilde\Psi(y,Q)=e^{ih_1Q}\psi(y),
\end{equation}
the function $\psi$ satisfying the second order differential
equation
\begin{equation}
\left[\partial_y^2-ih_1\partial_y-\left(h_2+\frac{g(g-1)}{\sin^2y}\right)\right]\psi
=0
\end{equation}
which, via the transformation $\psi(y)=\sin^g y\,\varphi(y)$,
can be rewritten as
\begin{equation}
\left[\partial_y^2+(2g\cot y-ih_1)\partial_y-(g^2+igh_1\cot y+h_2)\right]\varphi=0.
\end{equation}
The last equation, after the substitution $t=e^{2iy}$, is reduced
to the standard Fuchsian form
\begin{equation}\label{eq:Fuchs2}
\left[\partial_t^2+\left(
-\frac{g-1+\frac{1}{2} h_1}{t}+\frac{2g}{t-1}\right)\partial_t
+\left(
\frac{\frac{1}{4}(g^2+gh_1+h_2)}{t^2}-\frac{\frac{1}{2} gh_1}{t(t-1)}
\right)\right]\varphi=0.
\end{equation}
The equation (\ref{eq:Fuchs2}) has 3 regular singularities: $\{0,1,\infty\}$
with the characteristic exponents:
$$ \begin{array}{lll}
t\sim 1 & \varphi\sim(t-1)^\mu & \mu\in\{-2g+1,0\} \\
t\sim 0 & \varphi\sim t^\rho & \rho\in\{n_1,n_2+g\} \\
t\sim\infty & \varphi\sim t^{-\sigma} & -\sigma\in\{n_1-g,n_2\}
\end{array}
$$
Moreover, by the substitution $\varphi(t)\!=\!t^{n_1}(1-t)^{1-2g}f(t)$
the equation (\ref{eq:Fuchs2}) is reduced to the standard
hypergeometric equation
\begin{equation}
\left[t\partial_t(t\partial_t+b_1-1)-t(t\partial_t+a_1)(t\partial_t+a_2)\right]f=0,
\end{equation}
the parameters $a_1$, $a_2$, $b_1$ being given by the formulas
(\ref{eq:def-ab}) which for $N=2$ read
\begin{equation}\label{eq:def-ab2}
a_1=n_1-n_2+1-2g, \qquad
a_2=1-g, \qquad
b_1=n_1-n_2+1-g.
\end{equation}
{}From $J_{n_1n_2}\in{\cal T}^{(2)}$ it follows immediately that the
corresponding $\varphi_{n_1n_2}(t)$ is a Laurent polynomial in $t$.
\begin{prop}
The Laurent polynomial $\varphi_{n_1n_2}(t)$ is given, up to
a constant factor, by the formula (\ref{eq:phi-hgf}) which,
for $N=2$ takes the form
\begin{equation}\label{eq:phi-hgf2}
\varphi_{n_1n_2}(t)=t^{n_1}(1-t)^{1-2g}\,
{}_2F_1(a_1,a_2;b_1;t)
\end{equation}
the parameters $a_1$, $a_2$, $b_1$ being given by (\ref{eq:def-ab2}).
\end{prop}
{\bf Proof.}
Define the function $F_{n_1n_2}(t)$ by the right hand side of the
formula (\ref{eq:phi-hgf2}). Strictly speaking, the hypergeometric
series converges only for $|t|<1$ but in few moments we shall see
that $F_{n_1n_2}(t)$ continues analytically to the whole complex plane.
Using the well known formula
$$
(1-t)^{a+b-c}\,{}_2F_1(a,b;c;t)={}_2F_1(c-a,c-b;c;t)
$$
we can rewrite $F_{n_1n_2}(t)$ as folllows
$$
F_{n_1n_2}(t)=t^{n_1}\,{}_2F_1(n_1-n_2,g;n_1-n_2+1-g;t)
$$
It is easy to observe now that the hypergeometric series
in the right hand side terminates for integer $\{n_1\leq n_2\}$
and $F_{n_1n_2}$ is thus a Laurent polynomial
$$
F_{n_1n_2}=\sum_{k=n_1}^{n_2}t^k c_k(\vec n;g),
$$
of the form (\ref{eq:phi-Laurent}).
Since $F_{n_1n_2}$ satisfies the same differential equation
(\ref{eq:Fuchs2}) as
$\varphi_{n_1n_2}$ and the linearly independent solution to
(\ref{eq:Fuchs2}) is obviously not polynomial, the functions
$F_{n_1n_2}(t)$ and $\varphi_{n_1n_2}(t)$ are identical up to a
constant factor, which finishes the proof of the proposition
and of the conjectures 1 and 2 for $N=2$.\hfill\rule{2mm}{2mm}
\section{$A_2$ case: Integral transform}\label{int-tr}
For $N=3$ the commuting differential operators (\ref{eq:def-Hk})
read
\begin{eqnarray*}
H_1 & = & -i(\partial_1+\partial_2+\partial_3), \\
H_2 & = & -(\partial_1\partial_2+\partial_1\partial_3+\partial_2\partial_3)
-g(g-1)\left(\sin^{-2}q_{12}+\sin^{-2}q_{13}+\sin^{-2}q_{23}
\right), \\
H_3 & = & i\partial_1\partial_2\partial_3
+ig(g-1)\left(\sin^{-2}q_{23}\,\partial_1
+\sin^{-2}q_{13}\,\partial_2+\sin^{-2}q_{12}\,\partial_3
\right), \\
\end{eqnarray*}
and, respectively,
\begin{eqnarray*}
\tilde H_1&=&-i(\partial_1+\partial_2+\partial_3) \\
\tilde H_2&=&-(\partial_1\partial_2+\partial_1\partial_3+\partial_2\partial_3) \\
&&g{[}\cot q_{12}(\partial_1-\partial_2)+\cot q_{13}(\partial_1-\partial_3)
+\cot q_{23}(\partial_2-\partial_3){]} \\
&&-4g^2 \\
\tilde H_3&=& i\partial_1\partial_2\partial_3 \\
&&-ig{[}\cot q_{12}(\partial_1-\partial_2)\partial_3
+\cot q_{13}(\partial_1-\partial_3)\partial_2
+\cot q_{23}(\partial_2-\partial_3)\partial_1 {]}\\
&&+2ig^2{[} (1+\cot q_{12}\cot q_{13})\partial_1
+(1-\cot q_{12}\cot q_{23})\partial_2
+(1+\cot q_{13}\cot q_{23})\partial_3 {]}
\end{eqnarray*}
the vacuum function being
\begin{equation}
\Omega(\vec q)=\left|\sin q_{12}\sin q_{13}\sin q_{23}\right|^g.
\end{equation}
The eigenvectors $\Psi_{\vec n}$, resp.\ $J_{\vec n}$,
according to (\ref{eq:def-hm}), are parametrized
by the triplets of integers $\{n_1\leq n_2\leq n_3\}\in\hbox{\bbd Z}^3$,
the corresponding eigenvalues being
\begin{equation}\label{eq:def-h3}
h_1=2(m_1+m_2+m_3), \quad
h_2=4(m_1m_2+m_1m_3+m_2m_3), \quad
h_3=8m_1m_2m_3,
\end{equation}
where,
\begin{equation}\label{eq:def-m3}
m_1=n_1-g, \qquad
m_2=n_2, \qquad
m_3=n_3+g.
\end{equation}
The structure of the operator $K$ performing separation
of variables in the $A_2$ case is more complicated than
in the $A_1$ case. In contrast with the $A_1$ case, $K$ is
given by an integral operator rather then by simple change of
coordinates. To describe $K$, let us introduce the following notation.
$$ x_1=q_1-q_3, \qquad x_2=q_2-q_3, \qquad Q=q_3, $$
$$ x_\pm=x_1\pm x_2, \qquad y_\pm=y_1\pm y_2. $$
We shall study the action of $K$ locally, assuming
that $q_1>q_2>q_3$ and hence $x_+>x_-$.
The operator $K:\Psi(q_1,q_2,q_3)\mapsto\tilde\Psi(y_1,y_2;Q)$ is defined
as an integral operator
\begin{equation}\label{eq:def-K}
\tilde\Psi(y_1,y_2;Q)=\int_{y_-}^{y_+}d\xi\,{\cal K}(y_1,y_2;\xi)
\Psi\left(\frac{y_++\xi}{2}+Q,\frac{y_+-\xi}{2}+Q,Q\right)
\end{equation}
with the kernel
\begin{equation}\label{eq:kernel-K}
{\cal K}=\kappa\left[\frac{
\sin\left(\frac{\displaystyle \xi+y_-}{\displaystyle 2}\right)
\sin\left(\frac{\displaystyle \xi-y_-}{\displaystyle 2}\right)
\sin\left(\frac{\displaystyle y_++\xi}{\displaystyle 2}\right)
\sin\left(\frac{\displaystyle y_+-\xi}{\displaystyle 2}\right)}{
\sin y_1\sin y_2\sin \xi}\right]^{\displaystyle g-1}
\end{equation}
where $\kappa$ is a normalization coefficient to be fixed later.
It is assumed in (\ref{eq:def-K}) and (\ref{eq:kernel-K})
that $y_-<x_-=\xi<y_+=x_+$.
The integral converges when $g>0$ which will always be assumed
henceforth.
The motivation for such a choice of $K$ takes its origin
from considering the problem in the classical limit
$(g\rightarrow\infty)$ where there exists effective prescription
for constructing a separation of variables for an integrable system
from the poles of the so-called Baker-Akhiezer function.
See \cite{Skl:38}, \S7, for a detailed explanation.
\begin{theo}\label{Psi-diff-eq}
Let $H_k\Psi_{n_1n_2n_3}=h_k\Psi_{n_1n_2n_3}$.
Then the function $\tilde\Psi_{\vec n}=K\Psi_{\vec n}$
satisfies the differential equations
\begin{equation}
{\cal Q}\tilde\Psi_{\vec n}=0,\qquad
{\cal Y}_j\tilde\Psi_{\vec n}=0, \quad j=1,2
\end{equation}
where
\begin{equation}
{\cal Q}=-i\partial_Q-h_1,
\end{equation}
\begin{eqnarray}
\lefteqn{
{\cal Y}_j=i\partial_{y_j}^3+h_1\partial_{y_j}^2
-i\left(h_2+3\frac{g(g-1)}{\sin^2{y_j}}\right)\partial_{y_j}} \nonumber \\
&& -\left(h_3+\frac{g(g-1)}{\sin^2{y_j}}h_1
+2ig(g-1)(g-2)\frac{\cos {y_j}}{\sin^3 {y_j}}\right).
\end{eqnarray}
\end{theo}
The proof is based on the following proposition.
\begin{prop}\label{K:q-char-eq}
The kernel $K$ satisfies the differential equations
$$ [-i\partial_Q-H_1^*]K=0, $$
\begin{eqnarray*}
\lefteqn{ \left[i\partial^3_{y_j}+H_1^*\partial^2_{y_j}
-i\left(H_2^*+\frac{3g(g-1)}{\sin^2 y_j}\right)\partial_{y_j} \right.} \\
&& \left. -\left(H_3^*+H_1^*\frac{g(g-1)}{\sin^2 y_j}
+2ig(g-1)(g-2)\frac{\cos y_j}{\sin^3 y_j}\right)\right]K=0,
\end{eqnarray*}
where $H_n^*$ is the Lagrange adjoint of $H_n$
$$ \int\varphi(q)(H\psi)(q)\,dq=\int(H^*\varphi)(q)\psi(q)\,dq $$
\begin{eqnarray*}
H^*_1&=&i(\partial_{q_1}+\partial_{q_2}+\partial_{q_3}), \\
H^*_2&=&-\partial_{q_1}\partial_{q_2}-\partial_{q_1}\partial_{q_3}-\partial_{q_2}\partial_{q_3}
-g(g-1){[}\sin^{-2}q_{12}+\sin^{-2}q_{13}+\sin^{-2}q_{23}{]}, \\
H^*_3&=&-i\partial_{q_1}\partial_{q_2}\partial_{q_3}-ig(g-1)
{[}\sin^{-2}q_{23}\,\partial_{q_1}+\sin^{-2}q_{13}\,\partial_{q_2}
+\sin^{-2}q_{12}\partial_{q_3}{]}.
\end{eqnarray*}
\end{prop}
The proof is given by a direct, though tedious calculation.
To complete the proof of the theorem \ref{Psi-diff-eq}, consider
the expressions ${\cal Q}K\Psi_{\vec n}$ and ${\cal Y}_jK\Psi_{\vec n}$
using the formulas (\ref{eq:def-K}) and (\ref{eq:kernel-K}) for $K$.
The idea is to use
the fact that $\Psi_{\vec n}$ is an eigenfunction of
$H_k$ and replace $h_k\Psi_{\vec n}$ by $H_k\Psi_{\vec n}$.
After integration by parts in the variable $\xi$ the operators
$H_k$ are replaced by their adjoints $H_k^*$ and the result is zero
by virtue of proposition \ref{K:q-char-eq}.
The caution is needed however when handling the limits of integration
$y_\pm$ in (\ref{eq:def-K}). The following argument allows to
circumvent the problem of boundary terms. One can hide the limits
of integration into the definition of the kernel ${\cal K}$
considering the factors containing $(\xi-y_\pm)$ as the generalized
functions similar to $x_+^\lambda$, see \cite{G-Sh}. It is known
that $x_+^\lambda$ defined through the linear functional
$$ \left<f,x_+^\lambda\right>=\int_0^\infty dx\, f(x)x_+^\lambda
$$
is analytic in $\lambda$ on the complex plane excluding
the poles $x=-1,-2,\ldots$
and can be differentiated just as usual power function
$\partial_xx_+^\lambda=\lambda x_+^{\lambda-1}$. Therefore, we can safely
ignore the boundary of integral (\ref{eq:def-K}) while
integrating by parts. The only possible obstacle may
present the integer points $g=1,2,3$ (no more, since we need to
differentiate ${\cal K}$ maximum 3 times) where the boundary
may contribute delta-function terms. The direct calculation shows,
however, that all such terms cancel. \hfill\rule{2mm}{2mm}
The following theorem validates the conjectures 1 and 2 for the
$A_2$ case.
\begin{theo}\label{Psi3-fact}
The function $\tilde\Psi_{n_1n_2n_3}$ is factorized
\begin{equation}
\tilde\Psi_{n_1n_2n_3}(y_1,y_2;Q)=
e^{ih_1Q}\psi_{n_1n_2n_3}(y_1)\psi_{n_1n_2n_3}(y_2)
\end{equation}
according to (\ref{eq:factor-Psi}).
The separated function $\psi_{n_1n_2n_3}(y_2)$
has the structure (\ref{eq:fact-psi}).
\end{theo}
Note that, by virtue of the theorem \ref{Psi-diff-eq}, the function
$\tilde\Psi_{\vec n}(y_1,y_2;Q)$ satisfies an ordinary differential
equation in each variable. Since
${\cal Q}f=0$ is a first order differential equation
having a unique, up to a constant factor, solution $f(Q)=e^{ih_1Q}$,
the dependence on $Q$ is factorized. However, the differential
equations ${\cal Y}_j\psi(y_j)=0$ are of third order and have
three linearly independent solutions. To prove the theorem
\ref{Psi3-fact} one needs thus to study the ordinary
differential equation
\begin{eqnarray}
\lefteqn{
\left[ i\partial_y^3+h_1\partial_y^2
-i\left(h_2+3\frac{g(g-1)}{\sin^2y}\right)\partial_y\right.} \nonumber \\
&& -\left.\left(h_3+\frac{g(g-1)}{\sin^2y}h_1
+2ig(g-1)(g-2)\frac{\cos y}{\sin^3 y}\right)\right]\psi=0.
\label{eq:sep-eq3}
\end{eqnarray}
and to select its special solution corresponding to $\tilde\Psi$.
The proof will take several steps. First, let us eliminate from $\Psi$ and
$\tilde\Psi$ the vacuum factors $\Omega$, see (\ref{eq:Psi-Om}),
and, respectively
\begin{equation}\label{eq:tPsi-om}
\tilde\Psi(y_1,y_2;Q)=\omega(y_1)\omega(y_2)\tilde J(y_1,y_2;Q),
\qquad \omega(y)=\sin^{2g}y.
\end{equation}
Conjugating the operator $K$ with the vacuum factors
\begin{equation}\label{eq:conj-K}
M=\omega_1^{-1}\omega_2^{-1}K\Omega:J\mapsto\tilde J
\end{equation}
we obtain the integral operator
\begin{equation}\label{eq:def-M}
\tilde J(y_1,y_2;Q)=\int_{y_-}^{y_+}d\xi\,{\cal M}(y_1,y_2;\xi)
J\left(\frac{y_++\xi}{2}+Q,\frac{y_+-\xi}{2}+Q,Q\right)
\end{equation}
with the kernel
\begin{eqnarray}\label{eq:def-ker-M}
\lefteqn{
{\cal M}(y_1,y_2;\xi)={\cal K}(y_1,y_2;\xi)
\frac{\Omega\left(\frac{y_++\xi}{2}+Q,\frac{y_+-\xi}{2}+Q,Q\right)}%
{\omega(y_1)\omega(y_2)} } \nonumber \\
&&= \kappa\sin \xi\frac{
\left[
\sin\left(\frac{\xi+y_-}{2}\right)
\sin\left(\frac{\xi-y_-}{2}\right)
\right]^{g-1}
\left[
\sin\left(\frac{y_++\xi}{2}\right)
\sin\left(\frac{y_+-\xi}{2}\right)
\right]^{2g-1}}{\left[
\sin y_1\sin y_2\right]^{3g-1}}.
\end{eqnarray}
\begin{prop}\label{M:poly-poly}
Let $S$ be a trigonometric
polynomial in $q_j$, i.e.\ Laurent polynomial in $t_j=e^{2iq_j}$,
which is symmetric w.r.t.\ the transpositon $q_1\leftrightarrow q_2$.
Then $\tilde S=MS$ is a trigonometric polynomial symmetric w.r.t.\
$y_1\leftrightarrow y_2$.
\end{prop}
{\bf Proof.} It is more convenient to use variables $x_\pm$, $Q$
and, respectively, $y_\pm$, $Q$. Since the kernel ${\cal M}$
does not depend on $Q$ it is safe to omit the dependence on $Q$ in $S$.
The polynomiality and symmetry of $S$ are expressed now as
$S=S(x_+,x_-)=\sum_{k,n}s_{kn}e^{ikx_+}\cos nx_-$ where $k,n$ are
integers of the same parity, and $n\geq0$.
{}From (\ref{eq:def-M}),
(\ref{eq:def-ker-M}) we obtain
\begin{eqnarray*}
\lefteqn{\tilde S(y_+,y_-)=
\kappa\left(\sin^2\frac{y_+}{2}-\sin^2\frac{y_-}{2}\right)^{-3g+1}\times}\\
&&\times\int_{y_-}^{y_+}dx_-\,\sin x_-
\left(\sin^2\frac{x_-}{2}-\sin^2\frac{y_-}{2}\right)^{g-1}
\left(\sin^2\frac{y_+}{2}-\sin^2\frac{x_-}{2}\right)^{2g-1}
S(y_+,x_-).
\end{eqnarray*}
Let us make now the change of variables
\begin{equation}
\xi_\pm=\sin^2\frac{x_\pm}{2}, \quad
d\xi_\pm=\frac{1}{2}\sin x_\pm\,dx_\pm, \qquad
\eta_\pm=\sin^2\frac{y_\pm}{2},
\end{equation}
denoting $\check S(x_+,\xi_-)=S(x_+,x_-)$. It is easy to see that
$\check S(x_+,\xi_-)$ is polynomial in $\xi_-$ and that
\begin{equation}\label{eq:SS}
\tilde S(y_+,y_-)=
2\kappa(\eta_+-\eta_-)^{-3g+1}\int_{\eta_-}^{\eta_+}d\xi_-\,
(\xi_--\eta_-)^{g-1}(\eta_+-\xi_-)^{2g-1}\check S(y_+,\xi_-).
\end{equation}
Now put
$$ \xi_-=(\eta_+-\eta_-)\xi+\eta_- $$
and choose
\begin{equation}\label{def-kappa}
\kappa=\frac{1}{2B(g,2g)}=\frac{\Gamma(3g)}{2\Gamma(g)\Gamma(2g)}.
\end{equation}
Then, finally
\begin{equation}\label{eq:S-final}
\tilde S(y_+,y_-)=
\frac{\Gamma(3g)}{\Gamma(g)\Gamma(2g)}
\int_0^1 d\xi\, \xi^{g-1}(1-\xi)^{2g-1}
\check S(y_+,(\eta_+-\eta_-)\xi+\eta_-).
\end{equation}
It is sufficient to calculate the integral (\ref{eq:S-final})
for the monomials
$$ \check S=e^{iky_+}\eta_-^l(\eta_+-\eta_-)^m\xi^m $$
such that $k,l,m\in\hbox{\bbd Z}$, $l,m\geq 0$ and $k\equiv l+m \pmod{2}$.
Evaluating the beta-function integral
$$
\int_0^1 d\xi\, \xi^{g-1+m}(1-\xi)^{2g-1}=
\frac{\Gamma(g+m)\Gamma(2g)}{\Gamma(3g+m)}
$$
one obtains
\begin{equation}\label{eq:M-basis}
\tilde S(y_+,y_-)=\frac{\Gamma(3g)\Gamma(g+m)}{\Gamma(3g+m)\Gamma(g)}
e^{iky_+}\eta_-^l(\eta_+-\eta_-)^m.
\end{equation}
It is easy to verify that the result is a symmetric trigonometric
polynomial in $y_1$, $y_2$. \hfill\rule{2mm}{2mm}
Note that the normalization constant $\kappa$ is chosen in such a way
that $M:1\mapsto 1$.
The formula (\ref{eq:M-basis}) shows that the operator $M$
can in fact be continued analytically in $g$ on the whole complex
plane excluding the points $g=-\frac{1}{2},-1,-\frac{3}{2},\ldots$
coming from the poles of the gamma functions in (\ref{eq:M-basis})
and also
$g=-\frac{1}{3},-\frac{2}{3},\ldots$ coming from the poles
of $\Gamma(3g)$ in the normalization constant $\kappa$ (\ref{def-kappa}).
\section{$A_2$: Separated equation}\label{sep-eq}
To complete the proof of the theorem \ref{Psi3-fact}
we need to learn more about the separated equation (\ref{eq:sep-eq3}).
Eliminating from $\psi$ the vacuum factor $ \omega(y)=\sin^{2g}y $
via the substitution $\psi(y)=\varphi(y)\omega(y)$ one obtains
\begin{eqnarray}\label{eq:phi-eq-3}
\lefteqn{\left[i\partial_y^3+
(h_1+6ig\cot y)\partial_y^2\right.} \nonumber \\
&&+(-i(h_2+12g^2)+4gh_1\cot y+3ig(3g-1)\sin^{-2}y)\partial_y \nonumber \\
&&+\left.(-(h_3+4g^2h_1)-2ig(h_2+4g^2)\cot y
+g(3g-1)h_1\sin^{-2}y)\right]\varphi=0.
\label{eq:eq-phi3}
\end{eqnarray}
The change of variable $t=e^{2iy}$ brings the last equation to
the Fuchsian form:
\begin{equation}\label{eq:Fuchs3}
\left[\partial_t^3+w_1\partial_t^2+w_2\partial_t+w_3\right]\varphi=0
\end{equation}
where
\begin{eqnarray*}
w_1&\!\!=\!\!&-\frac{3(g-1)+\frac{1}{2} h_1}{t}
+\frac{6g}{t-1}, \\
w_2&\!\!=\!\!&\frac{(3g^2-3g+1)+\frac{1}{2}(2g-1)h_1+\frac{1}{4}h_2}{t^2}
+\frac{3g(3g-1)}{(t-1)^2}
-\frac{g(9(g-1)+2h_1)}{t(t-1)}, \\
w_3&\!\!=\!\!&-\frac{g^3+\frac{1}{2} g^2h_1+\frac{1}{4}gh_2+\frac{1}{8}h_3}{t^3}
+\frac{\frac{1}{2} g((h_2+4g^2)(t-1)-(3g-1)h_1)}{t^2(t-1)^2}.
\end{eqnarray*}
The points $t=0,1,\infty$ are regular singularities with the exponents
$$ \begin{array}{lll}
t\sim 1\quad & \varphi\sim(t-1)^\mu\quad & \mu\in\{-3g+2,-3g+1,0\} \\
t\sim 0\quad & \varphi\sim t^\rho\quad & \rho\in\{n_1,n_2+g,n_3+2g\} \\
t\sim\infty\quad & \varphi\sim t^{-\sigma}\quad & -\sigma\in\{n_1-2g,n_2-g,n_3\}
\end{array}
$$
Like in the $A_2$ case, the equation (\ref{eq:Fuchs3}) is reduced
by the substitution $\varphi(t)\!=\!t^{n_1}(1-t)^{1-3g}f(t)$
to the standard ${}_3F_2$ hypergeometric form \cite{Bateman}
\begin{equation}\label{eq:hge-32}
\left[t\partial_t(t\partial_t+b_1-1)(t\partial_t+b_2-1)
-t(t\partial_t+a_1)(t\partial_t+a_2)(t\partial_t+a_3)\right]f=0,
\end{equation}
the parameters $a_1$, $a_2$, $a_3$, $b_1$, $b_2$ being given by the formulas
(\ref{eq:def-hgf}) which for $N=3$ read
$$ a_1=n_1-n_3+1-3g, \qquad a_2=n_1-n_2+1-2g, \qquad
a_3=1-g,
$$
$$ b_1=n_1-n_3+1-2g, \qquad b_2=n_1-n_2+1-g. $$
\begin{prop}\label{Laurent-3}
Let the parameters $h_k$ be given by (\ref{eq:def-h3}),
(\ref{eq:def-m3}) for a
triplet of integers $\{n_1\leq n_2\leq n_3\}$ and $g\neq
1,0,-1,-2,\ldots$. Then the equation (\ref{eq:Fuchs3}) has
a unique, up to a constant factor, Laurent-polynomial solution
\begin{equation}
\varphi(t)=\sum_{k=n_1}^{n_3}t^k c_k(\vec n;g),
\end{equation}
the coefficients $c_k(\vec n;g)$ being rational functions of
$k$, $n_j$ and $g$.
\end{prop}
The above proposition follows from a more general statement.
\begin{theo}\label{Vadim}
Let the function $F_{n_1,\ldots,n_N}(t)$ be given for $|t|<1$ by
the right hand side of the formula (\ref{eq:phi-hgf}), the parameters
$a_j$ and $b_j$ being given by (\ref{eq:def-ab}) for some sequence
of integers $\vec n=\{n_1\leq n_2\leq\ldots \leq n_N\}$.
Let $g\neq 1,0,-1,-2,\ldots$. Then $F_{\vec n}(t)$ is a Laurent
polynomial
\begin{equation}
F_{\vec n}(t)=\sum_{k=n_1}^{n_N}t^k c_k(\vec n;g),
\end{equation}
the coefficients $c_k(\vec n;g)$ being rational functions of
$k$, $n_j$ and $g$.
\end{theo}
{\bf Proof.} Consider first the hypergeometric series (\ref{eq:def-hgf})
for ${}_NF_{N-1}$ which converges for $|t|<1$.
Using for $a_j$ and $b_j$ the
expressions (\ref{eq:def-ab}) one notes that
$a_{j+1}=b_j+n_{N-j+1}-n_{N-j}$ and therefore
$$
\frac{(a_{j+1})_k}{(b_j)_k}=
\frac{(b_j+k)_{n_{N-j+1}-n_{N-j}}}{(b_j)_{n_{N-j+1}-n_{N-j}}}.
$$
The expression
$$
\frac{(a_2)_k\ldots(a_N)_k}{(b_1)_k\ldots(b_{N-1})_k}=
\frac{(b_1+k)_{n_N-n_{N-1}}\ldots(b_{N-1}+k)_{n_2-n_1}}%
{(b_1)_{n_N-n_{N-1}}\ldots(b_{N-1})_{n_2-n_1}}=
P_{n_N-n_1}(k)
$$
is thus a polynomial in $k$ of degree $n_N-n_1$.
So we have
$$
{}_NF_{N-1}(a_1,\ldots,a_{N};b_1,\ldots,b_{N-1};t)=
\sum_{k=0}^\infty \frac{(a_1)_k}{k!}P_{n_N-n_1}(k)t^k
$$
from which it follows that
$$
{}_NF_{N-1}(a_1,\ldots,a_{N};b_1,\ldots,b_{N-1};t)=
\tilde P_{n_N-n_1}(t)(1-t)^{Ng-1}
$$
where $\tilde P_{n_N-n_1}(t)$ is a polynomial of degree $n_N-n_1$
in $t$. \hfill\rule{2mm}{2mm}
To prove now the proposition \ref{Laurent-3} it is sufficient to
notice that in the case $N=3$
the hypergeometric series ${}_3F_2(a_1,a_2,a_3;b_1,b_2;t)$
satisfies the same equation (\ref{eq:hge-32}) as $f(t)$ and
therefore the Laurent polynomial $F_{\vec n}(t)$ constructed
above satisfies the equation (\ref{eq:Fuchs3}).
The uniqueness follows from the fact that all the linearly
independent solutions to (\ref{eq:Fuchs3}) are nonpolynomial which
is seen from the characteristic exponents.
\hfill\rule{2mm}{2mm}
Now everything is ready to finish the proof of the theorem \ref{Psi3-fact}.
Since the function $\tilde J_{n_1n_2n_3}(y_1,y_2;Q)$ satisfies
(\ref{eq:phi-eq-3}) in variables $y_{1,2}$ and is a Laurent polynomial
it inevitably has the factorized form
\begin{equation}\label{eq:fact-J}
\tilde J_{n_1n_2n_3}(y_1,y_2;Q)=
e^{ih_1Q}\varphi_{n_1n_2n_3}(y_1)\varphi_{n_1n_2n_3}(y_2)
\end{equation}
by virtue of the proposition \ref{Laurent-3}. \hfill\rule{2mm}{2mm}
\section{Integral representation for Jack polynomials}\label{inv-K}
The formula (\ref{eq:fact-J}) presents an interesting
opportunity to construct a new integral representation of the
Jack polynomial $J_{\vec n}$ in terms of the ${}_3F_2$
hypergeometric polynomials $\varphi_{\vec n}(y)$ constructed above.
To achieve this goal, it is necessary to invert explicitely the
operator $M:J\mapsto\tilde J$.
Let us examine again the integral (\ref{eq:SS}).
Assume that $x_+=y_+$ and respectively $\xi_+=\eta_+$ are fixed
whereas $\xi_-$, $y_-$ are variables. Then, denoting
$$ \tilde s(\eta_-)=
\frac{1}{2\kappa\Gamma(g)}\tilde S(y_+,y_-)(\eta_+-\eta_-)^{3g-1},
\quad s(\xi_-)=(\eta_+-\xi_-)^{2g-1}\check S(y_+,\xi_-)
$$
we face the problem of inverting the integral transform
\begin{equation}\label{eq:frac-int}
\tilde s(\eta_-)=\int_{\eta_-}^{\eta_+}d\xi_-\,
\frac{(\xi_--\eta_-)^{g-1}}{\Gamma(g)}s(\xi_-)
\end{equation}
which is known as Riemann-Liouville integral of fractional order
$g$ \cite{Int-tr}. Its inversion is formally given by changing sign
of $g$
\begin{equation}
s(\xi_-)=\int_{\xi_-}^{\xi_+}d\eta_-\,
\frac{(\eta_--\xi_-)^{-g-1}}{\Gamma(-g)}\tilde s(\eta_-)
\end{equation}
and is called fractional differentiation operator.
However, by our assumption $g>0$, so the integrand becomes
singular at $\xi_-=\eta_-$ and the integral should be
regularized in the standard way \cite{G-Sh}.
Retracing all the intermediate transformations we
obtain
$$ S(x_+,x_-)=
\frac{\Gamma(2g)}{\Gamma(-g)\Gamma(3g)}
(\xi_+-\xi_-)^{-2g+1}\int_{\xi_-}^{\xi_+}d\eta_-\,
(\eta_--\xi_-)^{-g-1}(\xi_+-\eta_-)^{3g-1}
\tilde S(x_+,y_-)
$$
and finally come to the formula for $M^{-1}:\tilde J\mapsto J$
\begin{equation}\label{eq:JMJ}
J(x_+,x_-;Q)=\int_{x_-}^{x_+}dy_-\,
\check{\cal M}(x_+,x_-;y_-)\tilde J(x_+,y_-;Q)
\end{equation}
\begin{equation}\label{eq:inv-M}
\check{\cal M}=\check\kappa
\frac{\sin y_-
\left[
\sin\left(\frac{x_++y_-}{2}\right)
\sin\left(\frac{x_+-y_-}{2}\right)
\right]^{3g-1}}%
{\left[
\sin\left(\frac{y_-+x_-}{2}\right)
\sin\left(\frac{y_--x_-}{2}\right)
\right]^{g+1}
\left[\sin x_1\sin x_2\right]^{2g-1}}
\end{equation}
where
\begin{equation}
\check\kappa=\frac{\Gamma(2g)}{2\Gamma(-g)\Gamma(3g)}.
\end{equation}
For $K^{-1}$ we have respectively
\begin{equation}
\check{\cal K}=\check\kappa
\frac{\sin^g x_-\sin y_-
\left[
\sin\left(\frac{x_++y_-}{2}\right)
\sin\left(\frac{x_+-y_-}{2}\right)
\right]^{g-1}}%
{\left[
\sin\left(\frac{y_-+x_-}{2}\right)
\sin\left(\frac{y_--x_-}{2}\right)
\right]^{g+1}
\left[\sin x_1\sin x_2\right]^{g-1}}.
\end{equation}
The formulas (\ref{eq:fact-J}), (\ref{eq:JMJ}), (\ref{eq:inv-M})
provide a new integral representation for Jack polynomial
$J_{\vec n}$ in terms of the ${}_3F_2$
hypergeometric polynomials $\varphi_{\vec n}(y)$. The representation
would acquire more satisfactory form if one could describe
explicitely the normalization of $\varphi$ corresponding to the
standard normalization (\ref{eq:norm-J}) of $J$. We intend to study
this question in a subsequent paper.
It is remarkable that for positive integer $g$ the operators
$K^{-1}$, $M^{-1}$ become differential operators
of order $g$. In particular, for $g=1$ we have $K^{-1}=\partial/\partial y_-$.
\section{Separation of variables in the Schur polynomials}
\label{Schur}
For the generic $g$ the separation of variables
in Jack polynomials is so far unknown for $N>3$. However,
the problem simplifies drastically in the
case $g=1$, when Jack polynomials are reduced to the Schur polynomials
\cite{Macd}, and allows quite simple solution. In the present
section we have changed notation to make it more convenient
for handling Schur polynomials.
Let
\begin{equation}
P_{n_1\ldots n_N}(t_1,\ldots,t_N)=\det
\left|\begin{array}{cccc}
t_1^{n_1} & t_2^{n_1} & \ldots & t_N^{n_1} \\
t_1^{n_2} & t_2^{n_2} & \ldots & t_N^{n_2} \\
\ldots & \ldots & \ldots & \ldots \\
t_1^{n_N} & t_2^{n_N} & \ldots & t_N^{n_N}
\end{array}\right|.
\end{equation}
Schur polynomial is defined as the ratio of two antisymmetric
polynomials:
\begin{equation}
S_{\vec n}(\vec t)=\frac{P_{n_1,n_2+1,\ldots,n_N+N-1}(\vec t)}%
{P_{0,1,2,\ldots,N-1}(\vec t)}.
\end{equation}
Denominator (corresponding to $\Omega$ in the previous sections)
\begin{equation}
P_{0,1,2,\ldots,N-1}(\vec t)=\prod_{k>j}(t_k-t_j)
\end{equation}
is the elementary antisymmetric polynomial (Vandermonde determinant).
The separated equation
\begin{equation}
\prod_{j=1}^N \left(t\partial_t-n_j\right)\psi(t)=0.
\end{equation}
has as the general solution the polynomial
$ \psi(t)=\sum_{j=1}^N c_j t^{n_j}$.
The boundary condition
\begin{equation}
\left.\frac{\partial^k}{\partial t^k}\psi(t)\right|_{t=1}=0, \qquad
k=0,1,\ldots,N-2
\end{equation}
selects the solution
\begin{equation}
c_j\sim \det\left|\begin{array}{cccccc}
1 & \ldots & 1 & 1 & \ldots & 1 \\
n_1 & \ldots & n_{j-1} & n_{j+1} & \ldots & n_N \\
n_1^2 & \ldots & n_{j-1}^2 & n_{j+1}^2 & \ldots & n_N^2 \\
\ldots & \ldots & \ldots & \ldots & \ldots & \ldots \\
n_1^{N-2} & \ldots & n_{j-1}^{N-2} & n_{j+1}^{N-2} & \ldots & n_N^{N-2}
\end{array}\right|
=\prod_{\scriptstyle k>l \atop \scriptstyle k,l\neq j}(n_k-n_l).
\end{equation}
In case of Schur polynomials it is easier to construct
the inverse operator $K^{-1}$ rather than $K$. Let
\begin{equation}
\tilde\Psi(t_1,\ldots,t_{N-1})=\psi(t_1)\ldots\psi(t_{N-1})
=\prod_{j=1}^{N-1}\psi(t_j)
\end{equation}
and
\begin{equation}
K^{-1}=\prod_{k>j}\left(t_k\partial_{t_k}-t_j\partial_{t_j}\right).
\end{equation}
\begin{theo}
The operator $K^{-1}$ transforms the symmetric polynomial
$\tilde\Psi$ into an antisymmetric polynomial
$\Psi(t_1,\ldots,t_{N-1})=K^{-1}\tilde\Psi$
which is none other than the numerator of Schur polynomial
\begin{equation}
\Psi\left(\frac{t_1}{t_N},\ldots,\frac{t_{N-1}}{t_N}\right)
t^{n_1+\cdots+n_N}
\sim P_{n_1\ldots n_N}(t_1,\ldots,t_N).
\end{equation}
\end{theo}
The proof consists in an elementary calculation.
Since we have already seen in the $N=3$ case that $K^{-1}$ becomes
a differential operator for integer $g>0$,
it is not surprising that here $K^{-1}$ is also a differential operator.
\section{Discussion}
The construction of the operator $M$ performing the separation of
variables for Jack polynomials originates from mathematical physics
(Inverse Scattering Method) and contains a lot of guesswork.
A generalization of our results to the case of higher rank
$N>3$ could probably throw some light
on the algebraic and geometric meaning of the whole
construction which remains still obscure. The only available
results in this direction are so far the case $g=1$ (Schur
polynomials) and theorem \ref{Vadim} which allows to formulate
conjecture 2 about the structure of separated polynomials in the
general case.
Among other challenging problems one should mention generalizations
to other root systems, first of all $BC_N$, and also to the
$q$-finite-difference case (Macdonald polynomials).
\bigskip
\noindent{\it Acknowledgments.}
VK acknowledges support by the Nederlandse Organisatie voor
Wetenschappelijk Onderzoek (NWO).
ES wishes to thank
K.~Aomoto and S.G.~Gin\-di\-kin
for their interest in the work and valuable remarks.
|
\section{Introduction}
Dashen's theorem \cite{dash69} states that the squared mass
differences between the charged pseudoscalar mesons $\pi^\pm,
K^\pm$
and their corresponding neutral partners $\pi^0, K^0$ are equal in
the chiral limit, i.e., $\Delta M^2_K - \Delta M^2_\pi
= 0$, where $\Delta M^2_P = M^2_{P^\pm} - M^2_{P^0}$. In
recent years several groups have calculated the electromagnetic
corrections to this relation from non-vanishing quark masses. The
different conclusions are either that the violation is large
\cite{dono93,bij93} or that it {\it may be} large
\cite{mal90,ure95,neu94}.
\newline
The electromagnetic mass difference of the pions $\Delta M^2_\pi$
has
been determined in the chiral limit using current algebra by Das et
al. \cite{das67}. Ecker et al. \cite{eck89} have repeated the
calculation in the framework of chiral perturbation theory
($\chi$PT)
\cite{ga84} by resonance exchange within a photon loop. The
occurring divergences from these loops
are absorbed by introducing an electromagnetic counterterm (with a
coupling constant $\hat{C}$) in the chiral lagrangian. They find
that
the contribution from the loops is numerically very close to the
experimental mass difference, and thus conclude that the finite
part
of $\hat{C}$ is almost zero.\newline
In \cite{dono93} the authors have calculated the Compton scattering
of the
pseudoscalar mesons including the resonances and determined from
this amplitude the mass differences at order $O(e^2m_q)$. They concluded
first of all
by using three low-energy relations that the one-loop result is
finite,
i.e., there is no need of a counterterm lagrangian at order $O(e^2
m_q)$ in
order to renormalize the contributions from the resonances,
secondly they
found a strong violation of Dashen's theorem. We are in
disagreement with
both of these results.\newline
In this article we proceed in an analogous manner to \cite{eck89}
for the case
$m_q\neq 0$. We calculate in $\chi$PT the contributions of order
$O(e^2 m_q)$ to the masses of
the Goldstone bosons due to resonances. The divergences are
absorbed in the
corresponding electromagnetic counterterm lagrangian, associated
with the
couplings $\hat{K}_i$, where $i=1,\ldots,14$. The most general form
of
this
lagrangian has been given in \cite{ure95,neu94,bau95}. We find
again that
the contribution from the loops reproduces the measured mass
difference
$\Delta M^2_\pi$ very well,
and therefore we consider the finite parts of the $\hat{K}_i$
to be small. Using this assumption also for the calculation of
$\Delta M^2_K$, we may
finally read off the corrections to Dashen's theorem from one-loop
resonance exchange.
The (scale dependent) result shows that the resonances lead to rather
moderate deviations.\newline
The article is organized as follows. In section 2 we present the
ingredients from $\chi$PT and the resonances needed for
the calculation. In section 3 we give the contributions to the
masses
and to Dashen's theorem and renormalize the counterterm lagrangian.
The
numerical results and a short conclusion are given in section 4.
\pagestyle{plain}
\section{The Lagrangians at lowest and next-to-leading Order}
The chiral lagrangian can be expanded in derivatives of the
Goldstone
fields
and in the masses of the three light quarks. The power counting
is established in the following way: The Goldstone fields are of
order
$O(p^0)$, a derivative $\partial_\mu$, the vector and axial vector
currents
$v_\mu, a_\mu$ count as quantities of $O(p)$ and the scalar
(incorporating the masses) and pseudoscalar currents $s,p$ are of
order
$O(p^2)$. The effective lagrangian starts at $O(p^2)$,
denoted by ${\cal L}_2$. It is the non-linear $\sigma$-model
lagrangian
coupled
to
external fields, respects chiral symmetry $SU(3)_{R}\times
SU(3)_{L}$, and is invariant under $P$ and $C$ transformations
\cite{ga84},
\begin{eqnarray}\label{l2}
{\cal L}_2 &=&\frac{F^2_0}{4}\langle d^\mu U^\dagger d_\mu U + \chi
U^\dagger +
\chi^\dagger U\rangle\nonumber \\
d_{\mu} U &=& \partial_{\mu} U - i(v_{\mu} + a_{\mu}) U + i U
(v_{\mu} -
a_{\mu})\nonumber \\
v_{\mu}&=&QA_{\mu} + \cdots\nonumber \\
Q&=&\frac{e}{3}\;{\rm diag}\,(2,-1,-1)\nonumber \\
\chi&=&2B_0(s+ip)\nonumber \\
s&=&{\rm diag}\,(m_u,m_d,m_s)\nonumber \\
F_\pi&=&F_0 \left[1+O(m_q )\right]\nonumber \\
B_0&=&-\frac{1}{F^2_0}\langle 0|\bar{u}u|0\rangle\left[1+
O(m_q )\right] \quad .
\end{eqnarray}
The brackets $\langle\cdots\rangle$ denote the trace in flavour
space and
$U$ is a unitary $3\times 3$ matrix that incorporates the fields of
the
eight pseudoscalar mesons,
\begin{eqnarray}
U&=&\exp\,\left(\frac{i\Phi}{F_0}\right)\nonumber \\[4mm]
\Phi&=&\sqrt{2}\left(
\begin{array}{ccc}
\frac{1}{\sqrt{2}}\pi^0+\frac{1}{\sqrt{6}}\eta_8 & \pi^+ & K^+\\
\pi^- & -\frac{1}{\sqrt{2}}\pi^0+\frac{1}{\sqrt{6}}\eta_8 & K^0\\
K^- & \overline{K^0} & -\frac{2}{\sqrt{6}}\eta_8
\end{array}
\right) \quad .
\end{eqnarray}
Note that the photon field $A_{\mu}$ is incorporated in the vector
current $v_\mu$. The corresponding kinetic term has to be added to
${\cal L}_2$,
\begin{equation}
{\cal L}^\gamma_{kin} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu} -
\frac{1}{2}
\left(
\partial_\mu A^\mu \right)^2\, ,
\end{equation}
with $F_{\mu\nu}=\partial_\mu A_\nu -\partial_\nu A_\mu$ and the
gauge
fixing parameter chosen to be $\lambda =1$. In order to maintain
the usual
chiral counting in ${\cal L}_{kin}^\gamma$,
it is convenient to count the photon field as a
quantity of order $O(p^0)$, and the electromagnetic coupling $e$
of $O(p)$ \cite{ure95}.
\newline
The lowest order couplings of the pseudoscalar mesons to the
resonances are
linear in the resonance fields and start at order $O(p^2)$
\cite{eck89,eck289}. For the description of the fields we use the
antisymmetric tensor notation for the vector and axialvector
mesons, e.g., the vector octet has the form
\begin{equation}
V_{\mu\nu}=\left(
\begin{array}{ccc}
\frac{1}{\sqrt{2}}\rho^0_{\mu\nu}+
\frac{1}{\sqrt{6}}\omega_{8\,\mu\nu} &
\rho^+_{\mu\nu} & K^{\ast\,+}_{\mu\nu}\\[2mm]
\rho^-_{\mu\nu} &
-\frac{1}{\sqrt{2}}\rho^0_{\mu\nu}+
\frac{1}{\sqrt{6}}\omega_{8\,\mu\nu} &
K^{\ast\,0}_{\mu\nu} \\[2mm]
K^{\ast\,-}_{\mu\nu} & \overline{K^{\ast\,0}}_{\mu\nu} &
-\frac{2}{\sqrt{6}}\omega_{8\,\mu\nu}
\end{array}
\right) \quad .
\end{equation}
This method is discussed in detail in \cite{eck89}, we restrict
ourselves
on
the formulae needed for the calculations in the following
section. The relevant interaction lagrangian contains the octet
fields
only,
\begin{eqnarray}\label{lr}
{\cal L}^V_2&=&\frac{F_V}{2\sqrt{2}}\langle
V_{\mu\nu}f_+^{\mu\nu}\rangle
+\frac{iG_V}{2\sqrt{2}}\langle
V_{\mu\nu}[u^\mu,u^\nu]\rangle\nonumber \\
{\cal L}^A_2&=&\frac{F_A}{2\sqrt{2}}\langle
A_{\mu\nu}f_-^{\mu\nu}\rangle\nonumber \\
f_\pm^{\mu\nu}&=&u F^{\mu\nu}_L u^\dagger \pm u^\dagger
F^{\mu\nu}_R u\nonumber \\
F^{\mu\nu}_{R,L}&=&\partial^\mu (v^\nu\pm a^\nu)
-\partial^\nu(v^\mu\pm
a^\mu) - i[v^\mu\pm a^\mu,v^\nu\pm a^\nu]\nonumber \\
u^\mu&=&iu^\dagger d^\mu U u^\dagger =u^{\dagger\,\mu}\nonumber \\
U&=&u^2 \quad .
\end{eqnarray}
The coupling constants are real and are not restricted by chiral
symmetry
\cite{eck289}, numerical estimates are given in \cite{eck89}. In
the kinetic
lagrangian a covariant derivative acts on the vector and
axialvector mesons,
\begin{eqnarray}\label{rkin}
{\cal L}^R_{kin}&=&-\frac{1}{2}\langle \nabla^\mu R_{\mu\nu}
\nabla_\sigma R^{\sigma\nu} - \frac{1}{2}M^2_R
R_{\mu\nu}R^{\mu\nu}\rangle\hspace{2cm}R=V,A\nonumber \\
\nabla^\mu R_{\mu\nu}&=&\partial^\mu R_{\mu\nu} +
[\Gamma^\mu,R_{\mu\nu}]\nonumber \\
\Gamma^\mu&=& \frac{1}{2}\left\{u^\dagger [\partial^\mu
-i(v^\mu + a^\mu)]u + u[\partial^\mu -i(v^\mu - a^ \mu)]u^\dagger
\right\}\quad,
\end{eqnarray}
where $M_R$ is the corresponding mass in the chiral limit. Finally
we
collect all the different terms together into one lagrangian,
\begin{equation}\label{resonance}
{\cal L}^{eff}_2={\cal L}_2 + {\cal L}^R_2 + {\cal L}^\gamma_{kin}
+ {\cal
L}^R_{kin} \quad
{}.
\end{equation}
The one-loop electromagnetic mass shifts of the pseudoscalar mesons
calculated with this lagrangian (see section 3) contain divergences
that
can be absorbed in a counterterm lagrangian. In its general form,
this
lagrangian has one term of order $O(e^2)$ and 14 terms of
$O(e^2 p^2)$ \cite{ure95,neu94,bau95},
\begin{eqnarray}\label{l4}
{\cal L}^C_2&=&\hat{C}\langle Q U Q U^\dagger\rangle\nonumber \\
{\cal L}^C_4&=&\hat{K}_{1}F^2_0\langle d^\mu U^\dagger d_\mu
U\rangle
\langle Q^2\rangle
+\hat{K}_{2}F^2_0 \langle d^\mu U^\dagger d_\mu U\rangle
\langle Q U Q U^\dagger\rangle \nonumber \\
&&+\hat{K}_{3} F^2_0\left(\langle d^\mu U^\dagger Q U\rangle
\langle d_\mu U^\dagger Q U\rangle
+\langle d^\mu U Q U^\dagger\rangle \langle d_\mu U Q
U^\dagger\rangle
\right)\nonumber \\
&&+\hat{K}_{4} F^2_0 \langle d^\mu U^\dagger Q U\rangle
\langle d_\mu U Q U^\dagger \rangle \nonumber \\
&&+\hat{K}_{5} F^2_0 \langle \left( d^\mu U^\dagger d_\mu U
+ d^\mu U d_\mu U^\dagger\right) Q^2 \rangle\nonumber \\
&&+\hat{K}_{6} F^2_0 \langle d^\mu U^\dagger d_\mu U Q U^\dagger Q
U +
d^\mu U
d_\mu U^\dagger Q U Q U^\dagger\rangle\nonumber \\
&&+\hat{K}_{7} F^2_0\langle \chi U^\dagger + \chi^\dagger U\rangle
\langle Q^2\rangle
+\hat{K}_{8} F^2_0 \langle \chi U^\dagger + \chi^\dagger U \rangle
\langle Q U Q U^\dagger\rangle \nonumber \\
&&+\hat{K}_{9} F^2_0 \langle (\chi U^\dagger + \chi^\dagger U
+ U^\dagger \chi + U \chi^\dagger) Q^2\rangle\\
&&+\hat{K}_{10} F^2_0 \langle (\chi^\dagger U + U^\dagger \chi) Q
U^\dagger Q U
+ (\chi U^\dagger + U \chi^\dagger ) Q U Q
U^\dagger\rangle\nonumber \\
&&+\hat{K}_{11} F^2_0 \langle (\chi^\dagger U - U^\dagger \chi)Q
U^\dagger Q U
+ (\chi U^\dagger - U \chi^\dagger) Q U Q
U^\dagger\rangle\nonumber \\
&&+\hat{K}_{12} F^2_0 \langle d^\mu U^\dagger \left[ c^R_\mu Q , Q
\right] U
+ d^\mu U \left[ c^L_\mu Q , Q \right] U^\dagger\rangle\nonumber \\
&&+\hat{K}_{13} F^2_0 \langle c^{R\,\mu}Q U c^L_\mu Q
U^\dagger\rangle
+\hat{K}_{14} F^2_0 \langle c^{R\,\mu} Q c^R_\mu Q + c^{L\,\mu}
c^L_\mu
Q
\rangle\nonumber \\
&&+O(p^4,e^4)\nonumber \quad .
\end{eqnarray}
with $c^{R,L}_\mu Q = -i\left[v_\mu \pm a_\mu,Q\right]$.
The three last terms contribute only to matrix elements with
external
fields, we are therefore left with 12 relevant counterterms. Note
that we
have omitted terms which come either from the purely strong or
the purely electromagnetic sector in ${\cal L}^C_4$.\newline
At this point it is
worthwhile to discuss the connection of the present formalism to the usual
$\chi$PT without resonances where the Goldstone bosons and the (virtual)
photons are the only interacting particles. For this purpose we consider
the electromagnetic mass of the charged pion. In $\chi$PT the lagrangian
has the form up to and including $O(e^2p^2)$
\begin{eqnarray}
{\cal L} &=& {\cal L}_2^Q + {\cal L}_4^Q \nonumber \\
{\cal L}_2^Q &=& {\cal L}_2 + C\langle Q U Q U^\dagger\rangle,
\hspace{2cm}{\cal L}_4^Q = \sum^{14}_{i=1} K_i O_i
\end{eqnarray}
where $C$ and $K_i$ are low energy constants. They are independent of the
Goldstone bosons masses
and parameterize all the underlying physics (including
resonances) of
$\chi$PT. ${\cal L}_2$ is given in (\ref{l2}) and the operators $O_i$ are
identical to those in (\ref{l4}). Neglecting the
contributions of the order $O(e^2m_q)$ for a moment, the pion mass is
\cite{eck89}
\begin{equation}
M_{\pi^\pm}^2 = \frac{2e^2}{F_0^2}C + O(e^2m_q)
\end{equation}
entirely determined by the coupling constant $C$.
In the resonance approach $M_{\pi^\pm}^2$ gets contributions from
resonance-photon loops already at order $O(e^2)$ (see graphs (c)
and (d) in Figure 1)
\begin{equation}
M_{\pi^\pm}^2 = M_{\pi^\pm}^2|_{loops} + \frac{2e^2}{F_0^2}\hat C
+ O(e^2m_q)
\end{equation}
The loop term contains a divergent and a finite part and is completely
determined by the resonance parameters.
The divergences are absorbed by renormalizing
the coupling constants $\hat C$ \cite{eck89} (see section 3). The
connection to $\chi$PT without resonances is then given by the relation
\cite{eck89}
\begin{eqnarray}
C &=& C^R(\mu) + \hat C(\mu) \nonumber \\
C^R(\mu) &=& \frac{F_0^2}{2e^2}M_{\pi^\pm}^2|_{loops\;(finite)}
\label{finite}
\end{eqnarray}
where $C^R(\mu)$ and $\hat C(\mu)$ are finite and the scale dependence
cancels in the sum.
Relation (\ref{finite}) says that
the coupling constant $C$ is split in a part from resonances
$(C^R)$ and another part from non-resonant physics $(\hat C)$.
This ansatz of separating resonant and non-resonant contributions to the
low-energy parameters has been originally made for the strong interaction
sector at next to leading order \cite{eck89}. In this case
resonance
exchange gives tree-level contributions and no renormalization is
needed. In the electromagnetic case however,
contributions arise from resonances with photons in loops and we
renormalize the non-resonant part of the coupling constant, i.e. $\hat
C$ at order $O(e^2)$.\newline
In an analogous fashion the above procedure can be carried out up to
the order $O(e^2 p^2)$. The couplings $K_i$ of ${\cal L}_4^Q$ are in
general divergent, since they absorb the divergences of the one-loop
functional generated by ${\cal L}_2^Q$ \cite{ure95,neu94,bau95}.
At a specific scale point the renormalized coupling constants $K_i^r(\mu)$
can be split in two parts
\begin{equation}
K_i^r(\mu_0) = K_i^R(\mu_0) + \hat K_i(\mu_0)
\end{equation}
where the terms on the right-hand side are taken after renormalization
of $\hat K_i$ (see section 3) and are thus finite.
The choice of the scale point $\mu_0$ is not a priori fixed.
Like in the strong sector \cite{eck89} we consider $\mu_0$ in the range
of the lowest lying resonances, i.e. in the range from $0.5$ to $1.0
\mbox{ GeV}$.
In the strong sector it was found that
the resonances saturate the low-energy parameters
almost completely \cite{eck89}. In additon the authors have found
that the
same conclusion holds for the electromagnetic coupling constant
$C$ leading to $\hat C(\mu) \approx 0$.
Consequently we $assume$ that the $K_i^r(\mu)$ are also
saturated by resonance contributions, i.e. we put
\begin{equation}
\hat K_i(\mu) \approx 0 \quad.
\end{equation}
As we will see in section 4 this assumption works well in the case of
$\Delta M^2_\pi$.
\section{Corrections to Dashen's Theorem}
Using the lagrangian given in (\ref{resonance}) it is a
straightforward
process to
calculate the mass shift between the charged pseudoscalar mesons
$\pi^\pm,
K^\pm$ and their corresponding neutral partner $\pi^0, K^0$ at the
one-loop
level.
\begin{figure}[t]
\begin{center}
\begin{tabular}{c}
\epsfxsize=13.0cm
\leavevmode
\epsffile[67 358 546 746]{graphs.ps}
\end{tabular}
\caption[]{\label{fig1}One-loop contributions to the
electromagnetic mass
shift of $\pi^\pm$.}
\end{center}
\end{figure}
The relevant diagrams for the mass of the charged pion are shown in
Fig.1. Graph (a) contains
the off-shell pion form factor, (b) vanishes in
dimensional regularization and (c) is called ``modified seagull
graph''. Graph (d) contains an $a_1$-pole. The mass of the neutral
pion
does not get contributions from the loops.\newline
If we take the resonances to be in the $SU(3)$ limit according to
(\ref{rkin}), i.e., all vector resonances
have the same mass $M_V $ and all axialvector resonances the mass
$M_A$, we get the contributions listed below. For the graphs
with the pion form factor,
\begin{eqnarray}
\Delta_{p.f.}M^2_\pi &=& -ie^2 \int \frac{d^4 q}{(2\pi)^4}
\frac{q^2+4\nu+4M^2_\pi}{q^2(q^2+2\nu)}\nonumber
\\
&& -i\frac{8e^2 F_V G_V}{F^2_0} \int \frac{d^4q}{(2\pi)^4}
\frac{q^2M^2_\pi-\nu^2}{q^2 (q^2+2\nu) (M^2_V - q^2)}\\
&& -i\frac{4e^2 F^2_V G^2_V}{F^4_0}\int \frac{d^4 q}{(2\pi)^4}
\frac{q^2[q^2 M^2_\pi-\nu^2]}{q^2 (q^2+2\nu) (M^2_V - q^2)^2}
\nonumber \quad ,
\end{eqnarray}
where $\nu = pq$ and $p$ is the momentum of the pion. Using the
relation
$F_V G_V = F^2_0$
\cite{eck289} we obtain
\begin{equation}
\Delta_{p.f.}M^2_\pi = -ie^2 M^4_V \int \frac{d^4 q}{(2\pi)^4}
\frac{2\nu + 4M^2_\pi }{q^2(q^2+2\nu ) (M^2_V -
q^2)^2}
\quad .
\end{equation}
The modified seagull graph gives
\begin{equation}
\Delta_{s.g.}M^2_\pi = i \frac{e^2 F^2_V }{F^2_0} (3-\epsilon)
\int \frac{d^4 q}{(2\pi)^4}\frac{1}{M^2_V -
q^2}
\end{equation}
with $\epsilon = 4-d$, and finally for the $a_1$-pole graph, where
unlike
\cite{dono93} we get an additional second term,
\begin{eqnarray}\label{forgot}
\Delta_{a_1}M^2_\pi &=& -i \frac{e^2 F^2_A}{F^2_0}(3-\epsilon)
\int \frac{d^4
q}{(2\pi)^4}\frac{1}{M^2_A-q^2}
\nonumber \\
&-& i \frac{e^2 F^2_A}{F^2_0}
\int \frac{d^4 q}{(2\pi)^4}\frac{q^2\left[M^2_\pi
+(3-\epsilon)\nu \right]
+(2-\epsilon)\nu^2}{q^2\left[M^2_A-(q+p)^2\right]} \quad .
\end{eqnarray}
We now add the contribution from ${\cal L}^C_2$ and ${\cal L}^C_4$
to the mass shift
\cite{ure95,neu94} and evaluate the integrals,
\begin{eqnarray}\label{div}
\Delta M^2_\pi &=& -\frac{3e^2}{F^2_0 16\pi^2}
\left[ F^2_V M^2_V \left(\ln\frac{M^2_V}{\mu^2}
+ \frac{2}{3}\right)
-F^2_A M^2_A\left(\ln\frac{M^2_A}{\mu^2}
+ \frac{2}{3}\right)\right]\nonumber \\
&& -\frac{e^2 F^2_A}{F^2_0 16\pi^2} M^2_\pi
\left[2+\frac{3}{2}\ln\frac{M^2_A}{\mu^2}
+I_1\left(\frac{M^2_\pi}{M^2_A}\right)\right]\nonumber \\
&& +\frac{2e^2}{16\pi^2}M^2_\pi
\left[\frac{7}{2}-\frac{3}{2}\ln\frac{M^2_\pi}{M^2_V }
+I_2\left(\frac{M^2_\pi}{M^2_V}\right)\right] \\
&& +\frac{2e^2\hat{C}}{F^2_0} - \frac{6e^2}{F^2_0}
(F^2_V M^2_V - F^2_A M^2_A)\lambda\nonumber
\\
&& +8e^2M^2_K \hat{K}_{8}+2e^2M^2_\pi \hat{R}_{\pi}
-\frac{3e^2 F^2_A}{F^2_{0}}M^2_\pi\lambda\nonumber
\end{eqnarray}
with
\begin{eqnarray}
I_1(z) &=& \int^1_0 x\ln[x-x(1-x)z] \,dx\nonumber\\
I_2(z) &=& \int^1_0 (1+x)\left\{\ln[x+(1-x)^2 z]
-\frac{x}{x+(1-x)^2z}\right\}\,dx\nonumber \\
\hat{R}_\pi &=& -2 \hat{K}_3 + \hat{K}_4 + 2\hat{K}_8 +
4\hat{K}_{10} + 4\hat{K}_{11}
\end{eqnarray}
The divergences of the resonance-photon loops show up as poles
in $d=4$ dimensions. They are collected in the terms
proportional to $\lambda$
\begin{eqnarray}
\lambda &=& \frac{\mu^{d-4}}{16\pi^2} \left\{\frac{1}{d-4}
-\frac{1}{2}[\ln 4\pi+\Gamma'(1) +1 ]
\right\}\quad .
\end{eqnarray}
The occurring divergences are now canceled
by renormalizing the contributions from non-resonant physics, i.e. the
coupling constants $\hat C$ and $\hat K_{i}$. The divergence of the order
$O(e^2)$ (fourth
line of equation (\ref{div})) is absorbed by putting
\begin{eqnarray}
\hat C &=& \hat C(\mu)+3(F^{2}_{V}M^{2}_V-F^{2}_{A}M^{2}_{A})\lambda
\label{div1}
\end{eqnarray}
and that of the order $O(e^2 m_{q})$ (fifth line of equation
(\ref{div})) by the relation
\begin{eqnarray}
\hat R_{\pi} &=& \hat R_{\pi}(\mu)+\frac{3F^{2}_{A}}{2F^2_0}\lambda \quad.
\label{div2}
\end{eqnarray}
Using the second Weinberg sum rule \cite{wein67}
\begin{equation}\label{wein2}
F^2_V M^2_V - F^2_A M^2_A=0\quad ,
\end{equation}
the divergence in (\ref{div1}) cancels, but the divergence in
(\ref{div2}) does not. Even if we used an extension of this sum
rule to
order $O(m_q)$ \cite{pas82},
\begin{equation}
F^2_\rho M^2_\rho -F^2_{a_1}M^2_{a_1}\simeq F^2_\pi M^2_\pi\quad
\end{equation}
and assumed $F_A = F_0$ \cite{eck289}, the divergence would not
cancel,
on the contrary, it would become larger.\newline
We finally get the result
\begin{eqnarray} \label{finalpi}
\Delta M^2_\pi &=&-\frac{3e^2}{F^2_0 16\pi^2}
F^2_V M^2_V \ln\frac{M^2_V}{M^2_A}\nonumber\\
&& -\frac{e^2F^2_A}{F^2_0 16\pi^2} M^2_\pi
\left[2+\frac{3}{2}\ln\frac{M^2_A}{\mu^2}
+I_1\left(\frac{M^2_\pi}{M^2_A}\right)\right]\nonumber \\
&& +\frac{2e^2}{16\pi^2}M^2_\pi
\left[\frac{7}{2}-\frac{3}{2}\ln\frac{M^2_\pi }{M^2_V}
+I_2\left(\frac{M^2_\pi}{M^2_V}\right)\right]\\
&& +\frac{2e^2\hat{C}}{F^2_0}
+ 8e^2 M^2_K \hat{K}_8+2e^2 M^2_\pi \hat{R}_\pi
(\mu)\nonumber\quad ,
\end{eqnarray}
where we used (\ref{wein2}) to simplify the first term. In the
chiral limit
$\Delta M^2_\pi$ reduces to the expression given in
\cite{eck89}.\newline
The mass difference for the kaons is determined in an analogous
way, in the
contribution from the loops we merely have to replace $M^2_\pi$ by
$M^2_K$.
Finally the formula for the corrections to Dashen's theorem may be
read
off,
\begin{eqnarray}\label{dashen}
\Delta M^2_K - \Delta M^2_\pi &=&
- \frac{e^2 F^2_A}{F^2_0 16\pi^2}
\left\{M^2_K\left[2+\frac{3}{2}\ln\frac{M^2_A}{\mu^2}
+I_1\left(\frac{M^2_K}{M^2_A}\right)\right]\right.\nonumber
\\
&&\hspace{2.2cm}\left.- M^2_\pi\left[2
+\frac{3}{2}\ln\frac{M^2_A}{\mu^2}
+I_1\left(\frac{M^2_\pi}{M^2_A}\right)\right]\right\}\nonumber \\
&&+ \frac{2e^2}{16\pi^2}\left\{ M^2_K
\left[\frac{7}{2}-\frac{3}{2}\ln\frac{M^2_K}{M^2_V}
+I_2\left(\frac{M^2_K}{M^2_V}\right)\right]\right.\nonumber \\
&&\hspace{1.7cm}\left. - M^2_\pi
\left[\frac{7}{2}-\frac{3}{2}\ln\frac{M^2_\pi}{M^2_V}
+I_2\left(\frac{M^2_\pi}{M^2_V}\right)\right]\right\}\\
&&- 2e^2 M^2_K \left[\frac{2}{3}\hat{S}_K(\mu) + 4\hat{K}_8\right]
+ 2e^2 M^2_\pi\left[\frac{2}{3}\hat{S}_\pi -
\hat{R}_\pi(\mu)\right]\quad
,\nonumber
\end{eqnarray}
where $\hat{S}_{\pi,K}$ represent the contributions from the
counterterm
lagrangian to $\Delta M^2_K$,
\begin{eqnarray}
\hat{S}_\pi &=& 3\hat{K}_8 + \hat{K}_9 + \hat{K}_{10}\nonumber \\
\hat{S}_K &=& \hat{K}_5 + \hat{K}_6 - 6\hat{K}_8 - 6\hat{K}_{10} -
6\hat{K}_{11}\nonumber \\
\hat{S}_K &=& \hat{S}_K(\mu) + \frac{3F^2_A}{2F^2_0}\lambda
\quad .
\end{eqnarray}
\section{Numerical Results and Conclusion}
We put $F_0$ equal to the physical pion decay constant, $F_\pi =
92.4 \;{\rm MeV}$
and the masses of the mesons to $M_\pi = 135 \;{\rm MeV}, M_K = 495
\;{\rm MeV}$.
We take
$F_V = 154 \;{\rm MeV}$ \cite{eck89} and $M_V = M_\rho = 770 \;{\rm
MeV}$. To
eliminate the parameters of the axialvector resonances we use
Weinberg's
sum rules \cite{wein67},
\begin{equation}
F^2_V - F^2_A = F^2_0\hspace{2cm}
F^2_V M^2_V - F^2_A M^2_A=0 \quad .
\end{equation}
The contributions from the counterterm lagrangian are not known so far.
In \cite{eck89} it was found that the experimental mass difference
$\Delta M^2_\pi$ at order $O(e^2)$ is well reproduced by the
resonance-photon loops and therefore the authors conclude that the
contributions from non-resonant physics are small, i.e.
$\hat{C}\approx 0$. In analogy we $assume$ for the numerical
evaluation the
dominance of the resonant contributions at order $O(e^2m_q)$, i.e.
we put $\hat{K}_i(\mu)\approx 0$.\newline
Putting the numbers in
(\ref{finalpi}) we get for the contribution from the loops to
$\Delta M^2_\pi$
at the scale points $\mu=(0.5,0.77,1) \;{\rm GeV}$ (see Fig.2a)
\begin{equation}\label{numpi}
\Delta M^2_\pi |_{loops}= 2M_\pi\times (\;5.0\, ,\;5.1\, ,\;5.1\;)
\;{\rm MeV}\quad .
\end{equation}
which is in nice agreement with the experimental value $\Delta M^2_\pi
|_{exp.} = 2M_\pi\times 4.6\;{\rm MeV}$ \cite{pdg94}.
Using resonance saturation in the Kaon system as well,
we obtain for the corrections to Dashen's
theorem (again at
the scale points $\mu=(0.5,0.77,1) \;{\rm GeV}$)
\begin{equation}
\Delta M^2_K - \Delta M^2_\pi = (\;-0.13\, ,\;0.17\,
,\;0.36\;)\times10^{-3}\;({\rm GeV})^2
\quad ,
\end{equation}
which are smaller than the values found in the literature,
\begin{eqnarray}
\Delta M^2_K - \Delta M^2_\pi=\left\{
\begin{array}{ccc}
1.23&&\cite{dono93}\\
1.3\pm 0.4&\times 10^{-3} \;({\rm GeV})^2&\cite{bij93}\\
0.55\pm 0.25&&\cite{mal90}\\
\end{array}
\right.
\end{eqnarray}
\begin{figure}[t]
\begin{center}
\begin{tabular}{lr}
\epsfxsize=6.6cm
\leavevmode
\epsffile[18 144 593 718]{plots2a.ps} &
\epsfxsize=6.6cm
\leavevmode
\epsffile[18 144 593 718]{plots2b.ps}
\end{tabular}
\caption[]{\label{fig2} The solid lines show our results, the
dashed and
dotted curves represent in a) the experimental value \cite{pdg94},
in b)
the result of \cite{dono93}, respectively.}
\end{center}
\end{figure}
Of course, in order to get a scale independent result, the
counterterms are not allowed to vanish completely.
In \cite{dono93} the authors calculated the Compton scattering
of the
Goldstone bosons within the same model that
we have used in the present article and determined the corrections
to Dashen's theorem by closing the photon line. Their calculation
is finite
(without counterterms) and gives a considerably large value for
$\Delta M^2_K
- \Delta M^2_\pi$. The difference to our result may be
identified in (\ref{forgot}), where we have found an additional
(singular) term
that gives a large negative and scale dependent contribution. The
two
results are compared in
Fig.2b. Note that in \cite{dono93} the physical masses for
the
resonances
are used in the calculation of $\Delta M^2_K$, whereas we
work in the $SU(3)$ limit throughout.\newline
The other calculations are not strongly connected to our approach,
for a discussion of the value given in \cite{mal90} we refer to
\cite{ure95}.\newline
We therefore conclude that taking into account the resonances
at the one-loop level and working strictly in the $SU(3)$ limit for the
resonances leads to moderate rather than large corrections to Dashen's
theorem. Possibly strong violations must come from higher loop
corrections or from non-resonant physics.\\[8mm]
{\Large {\bf Acknowledgements}}\\[2mm]
We thank G.Ecker, J.Gasser, J.Kambor, H.Leutwyler and D.Wyler for
helpful
discussions.
|
\section{INTRODUCTION}
The remarkably successful application of the quark--parton model
in the description of deep-inelastic scattering (DIS) data over
a very large kinematic domain has propelled this
simple picture of the nucleon at high energies into becoming part
of the common language employed by medium and high energy physicists.
Furthermore, the QCD-improved parton model provides a framework
in which one can quantitatively understand the scaling violations
seen in DIS experiments in the perturbative region of large photon
virtualities ($Q^2 \agt 4$ GeV$^2$).
Nevertheless, what we have learned from the more recent DIS data
on both polarized and unpolarized targets is that even the QCD-improved
parton model cannot, by itself, give a complete description of the
structure of the nucleon at high energies.
It is unable to (nor was it intended to) explain the spectrum
of the nucleon's non-perturbative features.
Here one has traditionally invoked effective degrees of freedom,
for example in the form of a pionic cloud of the nucleon, to
describe the long range structure of the nucleon.
A very good example of this is the deviation from the QCD-parton
model prediction for the Gottfried sum rule \cite{GSR} seen in
the recent high-precision NMC data \cite{NMC}.
The most natural explanation of this result is that there exists
an excess of $\bar d$ quarks over $\bar u$ in the proton
--- something which is clearly impossible to obtain from
perturbative QCD alone.
A non-perturbative pionic cloud, on the other hand, offers
a simple explanation of this SU(2) flavor symmetry breaking
in the proton sea \cite{HM,SST,MTS,KL,MA,JUL}.
The more recent NA51 Drell-Yan experiment \cite{NA51} also strongly
suggests a suppression of the $\bar u$ sea in the proton
relative to the $\bar d$ sea.
Similarly in polarized DIS, the small value for the first moment
of the proton's spin-dependent structure function, $g_1$, obtained
initially by the EMC \cite{EMCG1}, and confirmed by later
measurements at CERN \cite{SMC} and SLAC \cite{SLACG1},
is widely interpreted as evidence of the breakdown of the simple
quark--parton model of nucleon structure.
The two most common interpretations of this result are
that either the strange sea of the proton is significantly
polarized, or that subtle anomaly effects (perhaps in the form
of highly polarized gluons) lead to a strong violation
of the OZI rule in the flavor singlet channel \cite{U1GLU}.
The simplest way to model a polarized strange sea would be in terms
of a polarized hyperon accompanying a non-perturbative cloud of kaons
\cite{TTT}.
DIS from a $\Lambda K$ or $\Sigma K$ component of the nucleon would
be a natural mechanism leading to a violation of the Ellis-Jaffe sum
rule \cite{EJ}.
While strongly suggesting that non-perturbative effects play an
important role in nucleon DIS, the results of these experiments
do not rule out mechanisms other than those involving meson clouds
as those responsible for the deviations of these sum rules from
the parton model predictions.
Indeed, despite the various phenomenological successes of nucleon
models which incorporate mesonic degrees of freedom, as yet there
is no direct experimental evidence unambiguously pointing to the
existence of a pion (or kaon) cloud in high energy reactions.
It is the purpose of this paper to identify experiments which could
give clear and unique signals of the presence of mesonic degrees
of freedom in nucleon DIS.
The role of pions in inclusive DIS from nucleon has been investigated
in a number of previous studies \cite{HM,SST,MTS,KL,MA,JUL,SUL,T83,FMS}.
Following Thomas \cite{T83}, it was realized that upper bounds
on the average pion number per nucleon could be extracted by
comparing with DIS data on the momentum fractions carried by
sea quarks in the proton.
Controversy as to whether all or just part of the Gottfried sum
rule violation can be accounted for by the pion cloud could be resolved
by obtaining a {\em lower} bound on the pion multiplicity.
However, to obtain a lower bound one would need to extract a
clear pionic signal from beneath the background arising from
the perturbative sea.
Since the pion contribution to the nucleon structure function
appears at relatively small Bjorken $x$ ($x \sim 0.1$), its signal
may be submerged beneath the perturbative background.
Therefore it seems a formidable challenge to seek direct
experimental confirmation of pionic effects in inclusive
DIS.
The problem is worse for the case of the $K$ cloud since, being
heavier, those contributions lie at even smaller $x$.
The pertinent question to ask is whether pions leave any unique
traces at all in other processes, which cannot be understood in
terms of perturbative quark and gluon degrees of freedom alone.
Recently in the literature several suggestions have been made
regarding the measurement of the pion cloud in other experiments.
Pirner and Povh \cite{HEID} have proposed to identify the size of
the constituent quark--pion vertex through exclusive leptoproduction
of fast pions in the current fragmentation region.
Dieperink and Pollock \cite{DP} have argued that one could obtain
information on the $\pi N$ form factor in DIS from a $^3He$ nucleus
by detecting the recoiling $^3He$ nucleus in the final state.
In the present paper we propose a series of analogous experiments
in semi-inclusive DIS on polarized protons, where a hadron is detected
in the final state in coincidence with the
scattered electron \cite{STA,LUS}.
We will demonstrate that DIS from the nucleon's pion cloud
(Fig.1) does in fact give rise to rather characteristic
fragmentation distributions in comparison with the predictions
of parton model hadronization.
These differences are significantly enhanced when initial
and final state polarization effects are considered.
We focus on semi-inclusive production of polarized $\Delta^{++}$
baryons from a polarized proton,
$e \vec p \rightarrow e' \vec\Delta^{++} X^-$.
Because the $g_1$ structure function of a pion is zero,
an unpolarized electron beam will suffice for this purpose.
The choice of the $\Delta^{++}$ for the final state baryon,
rather than, say, a nucleon, reduces the backgrounds that one
would have to consider due to the decay of $\Delta$s themselves.
Furthermore, the decay products of $\Delta^+$ or $\Delta^0$
would include neutral hadrons whose detection would be more
difficult, thus increasing the overall experimental uncertainties.
For the case of the $K$ cloud, the relevant reaction to observe
is $e \vec p \rightarrow e' \vec \Lambda X^+$.
Determining the polarization of the $\Lambda$ hyperon is
considerably easier because the $\Lambda$ is self-analyzing.
In the next Section we outline the basic kinematics
pertinent to semi-inclusive deep-inelastic scattering.
In Section III we present the predictions for the
polarization asymmetries in the pion cloud model
of the nucleon.
A condensed summary of the results in this Section can also
be found in Ref.\cite{CEBAF}.
Possible backgrounds to the pionic signal are
analyzed in Section IV.
In Section V the strangeness content of the nucleon is
studied in the $K$ cloud and diquark fragmentation models.
A brief overview of other experiments
suggested recently to measure the pion cloud of the nucleon
is given in Section VI, while Section VII is reserved for
some concluding remarks.
\section{KINEMATICS OF TARGET FRAGMENTATION}
Experimentally it is known that the yield of baryons is about
one order of magnitude higher in the backward hemisphere of the
$\gamma p$ center of mass frame (``target fragmentation region'')
than for forward hemisphere baryons (``current fragmentation region'')
\cite{BEB77,ARN85,EMC86}.
Furthermore, baryons produced by current fragmentation have
predominantly large momenta in the target rest frame
($\agt$ several GeV), while those in the backward center
of mass jet are generally slow.
Since our concern here is with low momentum baryons ($\Delta$s and
$\Lambda$s) produced in the target fragmentation region, we shall
neglect the quark $\rightarrow$ baryon fragmentation process which
gives rise to the forward baryons.
For studies of the spin dependence of the fragmentation process,
we require the target proton polarization to be parallel to the
photon direction, with the spin of the produced baryon quantized
along its direction of motion.
Experimentally, the polarization of the produced $\vec\Delta^{++}$
can be reconstructed from the angular distribution of its decay
products ($p$ and $\pi^+$), while because it is self-analyzing,
polarization of the $\Lambda$ can be determined automatically.
With the high luminosity beam available at CEBAF, for example,
the rate of $\Delta^{++}$ (or $\Lambda$) production will generally
be high.
Even though the efficiency with which low momentum baryons
can be accurately identified is lower than for fast baryons
in the forward center of mass hemisphere \cite{EMC86}, their
detection will still be feasible, for example with the CEBAF
Large Acceptance Spectrometer.
Alternatively, with polarized internal targets soon available
at HERMES, one could also in principle perform this experiment
there provided the $4\pi$ detectors will be capable of
capturing slow moving baryons as well as the fast mesons,
which will be the focus of the first stage of the HERMES
program.
Such a program would also be ideally suited for the proposed
European electron facility, ELFE, or the new Hadron Muon Collaboration
at CERN, which will be designed specifically for semi-inclusive
measurements.
We define our variables in the target rest frame as follows:
$l, l'$ are the four-momentum vectors of the initial and final
leptons;
$P_{\mu} = (M; 0, 0, 0)$ and
$p_{\mu} = (p_{0}; |{\bf p}| \sin\alpha \cos\phi,$
$|{\bf p}| \sin\alpha \sin\phi,$
$|{\bf p}| \cos\alpha)$
are the momentum vectors of the target proton and recoil baryon,
respectively; and
$q_{\mu} = (\nu; 0, 0, \sqrt{\nu^{2} + Q^{2}})$ denotes the photon
four-momentum, defined to lie along the positive $z$-axis.
Then $\nu = E-E'$ is the energy transferred to the target,
$y = \nu / E = 1 - E'/E$ is the fractional energy transfer
relative to the incident energy, and $Q^2 = -q^2 = 2 M E x y$\
is minus the four-momentum squared of the virtual photon, with
$x = Q^2 / 2 P \cdot q$.
With the possible CEBAF upgrade to $E \approx 8$--10 GeV, values of
$x \approx 0.13$--0.14 can be reached in the deep-inelastic region
for $\nu \approx 8$ GeV and $Q^2 \approx 2$ GeV$^2$,
corresponding to a center of mass energy squared of the
$\gamma p$ system of $W^2 = (P + q)^2 \sim 15$ GeV$^2$.
At HERMES, with a 30 GeV electron beam, one will comfortably probe
the $0.05 \alt x \alt 0.1$ region,
which is relevant for the pionic contribution,
up to $Q^2 \sim 5$ GeV$^2$ and $W^2 \alt 50$ GeV$^2$.
The four-momentum transfer squared between the proton and
baryon is
$t \equiv (P-p)^2 = -p^2_T / \zeta\ +\ t_{max}$,\
which is bounded from above
by $t_{max} = -(M_{B}^2 - M^2 \zeta) (1-\zeta)/\zeta$,
where $p_T^2 = {\bf p}^2 \sin^2\alpha$,
$\zeta = p \cdot q / P \cdot q$\ \ is the light-cone fraction
of the target proton's momentum carried by the baryon,
and $M_B$ is the recoiling baryon's mass.
In terms of $t$, the three-momentum of the produced baryon
is given by:
\begin{eqnarray}
|{\bf p}| &=& \frac{1}{2M} \sqrt{(M^2 + M_{B}^2 - t)^2
- 4 M^2 M_{B}^2}\ ,
\end{eqnarray}
so that in the target rest frame the slowest baryons are those
for which $t$ is maximized,
which occurs when $\zeta \rightarrow 1$.
As the upper limit on $\zeta$ is $1-x$, slow
baryon production also corresponds to the $x \rightarrow 0$ limit,
and the slowest possible particles produced at $\zeta = 1$
(at $x = 0$) will have momentum
$|{\bf p}_{min}| = (M_{B}^2 - M^2) / 2M \approx 340$ MeV
for $B=\Delta$,
and $\approx 193$ MeV for $B=\Lambda$.
For the pion-exchange process considered here, the peak in the
differential cross section occurs at $|{\bf p}| \sim 600$ MeV,
which, for $\zeta \sim 0.8$, corresponds to a missing mass of
$p_X^2 = (P-p+q)^2 \sim 0.8$ GeV$^2$
for $Q^2 \sim 2$ GeV$^2$ at CEBAF energies,
and
$p_X^2 \sim 5$ GeV$^2$
for $Q^2 \sim 4$ GeV$^2$ at HERMES.
In terms of the polar angle $\alpha$ (in the target rest frame),
\begin{eqnarray}
\cos\alpha
&=& \frac{ M_{B}^{2} + (1-2 \zeta) M^{2} - t }
{ \sqrt{(M_{B}^{2} - M^{2} - t)^{2} - 4 M^{2} t} },
\label{calpha}
\end{eqnarray}
between the $B$ and $\gamma$ momenta,
production of baryons will occur between $\alpha = 0$ and
\begin{eqnarray}
\alpha_{max}
&=& \arccos \left( \sqrt{1 - (M \zeta / M_{B})^2}
\right),
\end{eqnarray}
which for $\zeta \rightarrow 1$ is $\simeq 50^o$ for $B=\Delta$
and $\simeq 57^o$ for $B=\Lambda$.
For a given angle $\alpha$, the pion four-momentum will be
constrained to lie within the limits given by:
\begin{eqnarray}
t_{min/max}(\alpha)
&=& {1 \over \sin^2\alpha}
\left( M_{B}^2 \sin^2\alpha
- M^2 (1 - 2 x + \cos^2\alpha) \right.
\nonumber\\
&\pm&
\left. 2 M \cos\alpha\ \
\sqrt{ M^2 (1-x)^2
- M_{B}^2 \sin^2\alpha} \right).
\end{eqnarray}
At small angles baryons will be produced over essentially
the entire range of $t$
(and therefore $\zeta$), however the number will fall off rapidly as
$\alpha \rightarrow \arccos \left( \frac{1}{M_{B}}
\sqrt{M_{B}^2 - M^2 (1-x^2)}
\right)$
because of the fast convergence of the upper and lower bounds on $t$,
until no particles are produced beyond the kinematic boundary at
$t_{max} = t_{min}$
$ = - \left( M_{B}^2 (1 + x) - M^2 (1 - x) \right) / (1-x)$.
The importance of the above kinematic limits
was demonstrated in two experiments \cite{E745,BEBC}
in which slow proton production was studied in $\nu$-nucleon
and $\nu$-nucleus scattering.
The softening of the cross section for protons with
momentum less than ${\bf p}_{max}$ (equal to several hundred MeV in
the experiments), was shown \cite{IST,GRO,MTN} to be precisely due
to the absence of interactions at $x > x_{max}$, where\
$x_{max}
= 1 - (p_{0max} - |{\bf p}_{max}|)/M$.
The role of pions was also investigated in this process,
however due to the large perturbative sea component of the
nucleon structure function at $x \sim 0.05$, no definite
pionic signal could be identified.
We may hope, however, that by including polarization degrees
of freedom we can more efficiently isolate any pionic signal from
behind the fragmentation background.
\section{PION CLOUD DYNAMICS}
The pion model is a dynamical model of the nucleon where the
dissociation of a physical nucleon into a pion and an ``undressed''
nucleon or $\Delta$ is explicitly witnessed by the probing photon.
The possible relevance of the process illustrated in Fig.1, where
a $\pi^-$ emitted by the proton is hit by a photon, to DIS was
recognized some time ago \cite{SUL,STA,LUS}, and has since had several
important and interesting applications, most notably in providing
a mechanism to break SU(2) and SU(3) flavor symmetries in the
proton sea.
In the pion-exchange model the differential cross section is:
\begin{eqnarray}
{ d^5 \sigma \over dx dQ^2 d\zeta dp_T^2 d\phi }
&\propto&
{ f_{\pi N \Delta}^2 \over 16 \pi^2 m_{\pi}^2 }\
{ {\cal T}^{S\ s}(t)\ {\cal F}^2_{\pi \Delta}
\over (t - m_{\pi}^2)^2 }\
L_{\mu\nu}(l,q)\ W_{\pi}^{\mu\nu}(k,q),
\label{ope5}
\end{eqnarray}
where
$L_{\mu\nu} = 2 l'_{\mu} l_{\nu} + 2 l'_{\nu} l_{\mu}
- g_{\mu\nu} Q^2$
is the lepton tensor,
and
\begin{eqnarray}
W_{\pi}^{\mu\nu}
&=& - \left( g^{\mu\nu} + \frac{ q^{\mu} q^{\nu} }{ Q^2 } \right)\
W_{1\pi}\
+\ \left( k^{\mu} + \frac{ k \cdot q }{ Q^2 } q^{\mu} \right)
\left( k^{\nu} + \frac{ k \cdot q }{ Q^2 } q^{\nu} \right)\
\frac{ W_{2\pi} }{ m_{\pi}^{2} },
\label{Wpi}
\end{eqnarray}
describes the $\gamma \pi$ vertex, with $k$ denoting the virtual
pion four-momentum.
The quantity $T^{S\ s}(t)$ is the amplitude for a nucleon of
spin $S$ to emit a pion of four-momentum squared $t$, leaving
a $\Delta$ with spin $s$.
Since in the final analysis we will be dealing with Lorentz-invariant
cross sections as a function of the Lorentz-scalars $x$ and $\zeta$,
we can, without loss of generality, formulate the problem in any
frame which will simplify the analysis.
Here we note that factorization of the $\gamma N$ cross section
into $\gamma\ \pi$ and $\pi\ N$ (or $\gamma\ N^*$ and $N^* N$)
cross sections does not hold in all frames of reference \cite{MST}.
Indeed, such factorization, or convolution, can only be achieved
by eliminating antiparticle degrees of freedom, which can
formally be done only in the infinite momentum frame (IMF)
or on the light-cone \cite{WEIN,DLY}.
Therefore for the $\pi N \Delta$ form factor in Eq.(\ref{ope5})
we take the form suggested in earlier IMF studies of the
pionic content of the proton in inclusive DIS \cite{MTV}:
\begin{mathletters}
\begin{eqnarray}
\label{FF}
{\cal F}_{\pi \Delta}(p_T^2,\zeta) &=&
\left( { \Lambda^2 + M^2 \over
\Lambda^2 + s_{\pi \Delta} }
\right)^2,
\end{eqnarray}
where $s_{\pi \Delta} \equiv (p + k)^2
= (m_{\pi}^2 + p_T^2)/(1-\zeta)
+ (M_{\Delta}^2 + p_T^2)/\zeta$.
Since the form factor in the IMF is not yet very well
constrained, other forms for its shape \cite{ZOL} are also
possible (it has been suggested in Ref.\cite{ZOL} to use
semi-inclusive $NN$ scattering data as a means of obtaining
an upper bound on $\Lambda$, although here one also has to
deal with contributions from competing Reggeized meson
exchanges \cite{AG}).
However, the precise shape of the form factor is not
important here, since, as we shall see, the bulk of the effect
is given entirely by the proton--pion spin correlations.
Indeed, covariant formulations with $t$-dependent form factors
\cite{HM,SST,MTS,KL,MA,JUL,MSM}:
\begin{eqnarray}
\label{FFt}
{\cal F}_{\pi \Delta}(p_T^2,\zeta) &=&
\left( { \Lambda^2 - M^2 \over
\Lambda^2 - t(p_T^2,\zeta) }
\right)^2,
\end{eqnarray}
\end{mathletters}
give very similar results to those with the
$s_{\pi\Delta}$-dependent forms in Eq.(\ref{FF}).
The formulation in the IMF also allows one to use the on-mass-shell
structure function of the pion in Eq.(\ref{Wpi}) \cite{MTV,ZOL,GCPP},
without the need to model the extrapolation of the off-shell pion
structure function into the $t \not= m_{\pi}^2$ region
\cite{PISF,SHAKIN}.
(Although, in principle, there could be effects in the virtual pion
structure function due to the off-energy-shell dependence.)
For the pion structure function we use therefore the most
recent parametrization \cite{SMRS} of data extracted from
Drell-Yan experiments \cite{NA10}.
The main uncertainty in the covariant calculation is in fact the
off-mass-shell extrapolation of the virtual pion structure function,
for which there still does not exist consensus in the literature
\cite{PISF,SHAKIN}.
The $\pi N \Delta$ coupling constant, $f_{\pi N \Delta}$, in
Eq.(\ref{ope5}) is the physical coupling constant, defined
at the pion pole ($t = m_{\pi}^2$).
Note that there is no renormalization factor, $Z$,
multiplying the pion-exchange cross section, as has been
used recently in Refs.\cite{JUL,SB}.
This factor, which to first order in $f_{\pi N \Delta}$ is
written $Z = 1 / (1 + < n >_{\pi\Delta})$, with the pion number
$< n >_{\pi\Delta}$ being essentially the integrated cross section
in Eq.(\ref{ope5}), is usually introduced to normalize the total
nucleon inclusive cross section in the presence of pions
\cite{SST,MTV}.
For the physical, semi-inclusive process, however, its use
would lead to an artificial suppression of the pion-exchange
contribution, especially when the form factor is hard.
The authors of Ref.\cite{SB} also use convolution formulae
within a covariant framework, which, as mentioned above,
inherently makes use of the assumption of factorization as well
as the $k^2$-independence of the off-shell pion structure function,
the justification of which has not yet been demonstrated.
The function ${\cal T}^{S\ s}(t)$ in Eq.(\ref{ope5}) is obtained
by evaluating the trace over the target nucleon spinor and the
Rarita-Schwinger spinor-vector $u_{\alpha}$ for the recoil $\Delta$:
\begin{eqnarray}
{\cal T}^{S\ s}(t)
&=& {\rm Tr} \left[ u(P,S) \bar u(P,S)\
u_{\alpha}(p,s) \bar u_{\beta}(p,s)
\right]
(P - p)^{\alpha} (P - p)^{\beta},
\end{eqnarray}
where \cite{RS}
\begin{mathletters}
\begin{eqnarray}
u_{\alpha}(p,s)
&=& \sum_m \left\langle
{3 \over 2}\ s \left| 1\ m; {1 \over 2}\ s-m \right.
\right\rangle
\epsilon_{\alpha} (m)\ u(p,s-m)
\end{eqnarray}
is constructed from the spin-1/2 Dirac spinor $u$ and spin-1
vectors $\epsilon_{\alpha}(m)$, and normalized such that \cite{BDM}:
\begin{eqnarray}
\sum_s u_{\alpha}(p,s) \bar u_{\beta}(p,s)
&=& \Lambda_{\alpha \beta}(p),
\end{eqnarray}
\begin{eqnarray}
\Lambda_{\alpha \beta}(p)
&=& (\not\!p + M_{\Delta})
\left( - g_{\alpha\beta}
+ {\gamma_{\alpha} \gamma_{\beta} \over 3}
+ {\gamma_{\alpha} p_{\beta} - \gamma_{\beta} p_{\alpha}
\over 3\ M_{\Delta}}
+ {2\ p_{\alpha} p_{\beta} \over 3\ M_{\Delta}^2}
\right).
\end{eqnarray}
\end{mathletters}%
Because it is emitted collinearly with the pion, production of
$\Delta$ baryons with helicity $\pm 3/2$ is forbidden,
which leads to the selection rule:
\begin{eqnarray}
{\cal T}^{S\ \pm {3 \over 2}}(t) &=& 0.
\end{eqnarray}
This is confirmed by explicit evaluation of the trace
if we recall that for polarized fermion spinors the spin
projection is
$u(P,S) \overline{u}(P,S) = (1 + \gamma_5 \not\!\!\!S)
(\not\!\!\!P + M)/2$.
The polarization vectors $S$ and $s$ can be parametrized as:
$S = (0;0,0,+1)$ and
$s = \pm
\left.
\left(\sqrt{p_0^2-M_B^2}; p_0 \sin\alpha \cos\phi,
p_0 \sin\alpha \sin\phi,
p_0 \cos\alpha \right) \right/ M_B$,
so that the angle $\alpha$ between the polarization vectors of the
target proton $S$ and recoiling baryon $s$ coincides with the
direction of the momentum vector ${\bf p}$ relative to the
$z$-axis.
The yield of spin projection $\pm 1/2$\ states is then given by:
\begin{eqnarray}
{\cal T}^{ +{1 \over 2}\ \pm{1 \over 2} }(t)
&=& { 1 \over 12 M_{\Delta}^2 }
\left[ (M - M_{\Delta})^2 - t \right]\
\left[ (M + M_{\Delta})^2 - t \right]^2\
(1 \pm \cos\alpha).
\end{eqnarray}
Because the production of $\Delta$ baryons is limited to
forward angles in the target rest frame,
the factor $(1 \pm \cos\alpha)$
associated with the final state polarization will significantly
suppress the $s = -1/2$ yield relative to that of $s = +1/2$ final
states.
The differential cross section, $Q^2 d^3\sigma / dx dQ^2 d\zeta$,
for the individual polarization states of the produced $\Delta^{++}$
(for DIS from a proton with $S=+1/2$) is shown in Fig.2a for
typical CEBAF kinematics,
$x = 0.14$, $Q^2 = 2$ GeV$^2$ and $E = 8$ GeV, and in Fig.2b for
$x = 0.075$, $Q^2 = 4$ GeV$^2$ and $E = 30$ GeV, as may be expected
at HERMES.
The pion-exchange model predictions (solid curves)
use the form factor in Eq.(\ref{FF}) with cut-offs
$\Lambda =$ 600 (smallest), 800 and 1000 (largest) MeV,
which gives $< n >_{\pi\Delta} \approx$ 0.01, 0.02 and 0.04,
respectively.
(For $< n >_{\pi\Delta} \approx 0.02$ the cut-off in a $t$-dependent
dipole form factor would be $\sim 700$ MeV.)
The spectrum shows strong correlations between the polarizations
of the target proton ($S=+1/2$) and the $\Delta^{++}$.
In the next Section we examine the extent to which the suppression
of the antiparallel configuration of the $p$ and $\Delta$ spins
in the pion-exchange model is diluted by the competing parton
fragmentation process, which constitutes the main background to the
pion-exchange process discussed here.
\section{BACKGROUNDS}
At the large energy and momentum transfers possible with high-energy
($E \agt 10$ GeV) electron beams, the resonance backgrounds
should not pose a major problem in identifying the required signal.
Firstly, interference from quasi-elastic $\Delta^{++}$ production
will be eliminated by charge conservation.
Secondly, the large $W$ involved means that interference from
excited $\Delta^*$ states (with subsequent decay to $\Delta^{++}$
and pions) will be negligible.
In addition, any such resonance contributions will be strongly
suppressed by electromagnetic form factors at large $Q^2$
($Q^2 \agt 2$ GeV$^2$).
A potentially more significant background will be that due to
uncorrelated spectator fragmentation, as illustrated in Fig.3.
We can estimate the importance of this process within the parton
model framework, in which the cross section is proportional to
(assuming factorization of the $x$ and $\zeta$ dependence
\cite{FF,SSV,SCHM}):
\begin{eqnarray}
\frac{ d^4\sigma^{(s)} }{ dx dQ^2 dz dp_T^2 }
&\propto& {\cal F}_{p\uparrow}(x,Q^2)\
\widetilde{D}_{p\uparrow-q\uparrow\downarrow}^{s}(z,p_T^2),
\end{eqnarray}
where $z = \zeta/(1-x)$ is the light-cone momentum fraction
of the produced baryon carried by the spectator system.
The function ${\cal F}_{p\uparrow}(x,Q^2)$ is proportional to the
spin-weighted interacting-quark momentum distribution functions,
$q^{\uparrow\downarrow} (x) = (q(x) \pm \Delta q(x) )/2$,
where $^{\uparrow \downarrow}$ denote quark spins
parallel or antiparallel to the spin of the proton,
with $q(x)$ and $\Delta q(x)$ being the sum and difference
of $q^{\uparrow}$ and $q^{\downarrow}$, respectively.
For our numerical estimates we use the parametrization of
$\Delta q(x)$ from Gehrmann and Stirling \cite{GS},
and the CTEQ \cite{CTEQ} parametrization for $q(x)$.
The results change little if one uses, for example, the models
of Carlitz and Kaur \cite{CK} or
Sch\"afer \cite{SCHAEFER} for $\Delta q(x)$.
The fragmentation function
$\widetilde{D}_{p\uparrow-q\uparrow\downarrow}^{s}(z,p_T^2)$ gives
the probability for the polarized
($p^{\uparrow}$ minus $q^{\uparrow\downarrow}$)
spectator system to fragment into a $\Delta^{++}$ with
polarization $s$.
The usual assumption is that the transverse momentum distribution
of the baryon also factorizes \cite{SSV,RW,BRE,ACK},\
$\widetilde{D}_{p\uparrow - q\uparrow\downarrow}^s(z,p_T^2)
= D_{p\uparrow - q\uparrow\downarrow}^s(z)\ \varphi(p_T^2)$,
with $\int dp_T^2\ \varphi(p_T^2) = 1$.
To describe the soft, non-perturbative parton fragmentation process,
a number of phenomenological models have been developed for the
fragmentation functions.
Many of these \cite{SLO,BFM} have followed the basic approach
originally formulated by Field and Feynman \cite{FF,FFFRAG},
whose quark jet fragmentation model involved recursive $q \bar{q}$
pair creation (cascade) out of the color field between the scattered
and spectator partons, with subsequent recombination into color
neutral hadrons.
In the original analysis of Ref.\cite{FF} only (unpolarized)
quark $\rightarrow$ meson fragmentation functions were modeled.
Later this approach was extended by Sukhatme et al. \cite{SLO},
and Bartl et al. \cite{BFM} by also allowing for $q \rightarrow$
baryon and $qq \rightarrow$ baryon decays.
The approach pioneered by the Lund group \cite{LUND} included, in
addition, the fragmentation into hadrons of the gluon string
connecting the colored partons.
Analytic expressions for the fragmentation functions can be obtained
by constraining their limiting behavior at the asymptotic limits.
The $z \rightarrow 0$ limit requires a $1/z$ behavior for $D(z)$ in
order to reproduce the observed logarithmic increase in hadron
multiplicity as $s \rightarrow \infty$,
\begin{eqnarray}
<N_B> &=& \int_{z_{min}}^1\ dz\ D(z)\ \ \sim\ \ \ln s,
\label{nB}
\end{eqnarray}
where $z_{min} \propto 1/s$ (see below).
For the $z \rightarrow 1$ limit one commonly applies dimensional
counting rules \cite{FPS}, using essentially the same arguments
as for the $x \rightarrow 1$ limit of structure functions \cite{COUNT}.
For the specific case of the $\Delta^{++}$, at large $z$ this should
carry most of the parent system's momentum, and therefore contain
both valence $u$ quarks from the target proton.
In our region of interest, namely $z \agt 0.6$, where the pionic
contribution is the largest, by far the most important contributions
to $D(z)$ come from the process whereby the $\Delta^{++}$ is formed
after only one $u \bar{u}$ pair is created \cite{SLO,BFM}.
As a consequence, DIS from valence $u$ quarks will not be too
important.
For scattering from sea quarks we assume the same fragmentation
probabilities for $uuq\bar q$ spectator states as for $uu$,
although in general multi-quark configurations could decay at
different rates than the valence diquark
(however, already at $Q^2 \simeq 2$ GeV$^2$ the sea constitutes
at most $\sim 15\%$ of the cross section at $x \sim 0.1$).
Rather than rely on model counting rule arguments,
we parametrize the (very limited) EMC data \cite{EMC86}
on unpolarized $\Delta^{++}$ muon production
for $z \rightarrow 1$ as:
$D_{uu}(z \rightarrow 1) = \alpha (1 - z)^{\beta}$,
where $\beta \approx 0.3$.
The overall normalization of the fragmentation function
is fixed by the data to be
$\alpha \approx 0.68$.
Note that in obtaining this parametrization it has been necessary
to perform a conversion of the kinematic variables.
Usually in semi-inclusive experiments \cite{EMC86} the longitudinal
momentum dependence is measured as a function of the Feynman variable
$x_F$, defined as the ratio of the center of mass longitudinal
momentum to its maximum allowed
value, $x_F = p^*_L / p^*_{L max}
\simeq 2 |{\bf p}^*| / \sqrt{W^2}
\simeq 1 - M_X^2 / W^2$,
where $M_X$ is the mass of the inclusive hadronic debris,
and the asterisk ($^*$) denotes center of mass momenta.
This variable can be related to the light-cone variable $z$ via
\begin{eqnarray}
z &=& { \sqrt{ M_{\Delta}^2 + p_T^2 + W^2\ x_F^2/4 }
- \sqrt{W^2}\ x_F/2
\over \sqrt{W^2} }.
\end{eqnarray}
Note that for $z \rightarrow 1$, $x_F \simeq z$
if $W^2 \gg M_{\Delta}^2 + p_T^2$.
The target (current) fragmentation region corresponds to
$x_F < 0$ ($x_F > 0$), and the boundary between the regions
at $x_F = 0$ corresponds to
$\zeta_{min} = \sqrt{M_{\Delta}^2 + p_T^2} / \sqrt{W^2}$.
To model the spin dependence of the fragmentation process we follow
the simple approach taken by Bartl et al. \cite{BFMS} (see also
Refs.\cite{BIGI,DON}) in their study of polarized quark $\rightarrow$
baryon fragmentation.
Namely, the diquark is assumed to retain its helicity
during its decay, and the $q \bar{q}$ pair creation probability
is independent of the helicity state of the quark $q$.
At leading order this means that the produced baryon contains
the helicity of the diquark, so that, for example,
a $\Delta^{\Uparrow}$ or $\Delta^{\uparrow}$ can
emerge from a $q^{\uparrow} q^{\uparrow}$ diquark,
whereas a $\Delta^{\Downarrow}$ cannot.
(Our notation here is that
$\Uparrow, \uparrow, \downarrow, \Downarrow$
represent $s = +3/2, +1/2, -1/2, -3/2$ states, respectively.)
The overall normalization of the spin-dependent fragmentation
functions is fixed by the condition
\begin{eqnarray}
q(x)\ D_{p-q}(z)\
+\ \bar{q}(x)\ D_{p-\bar{q}}(z)\
&=& q^{\uparrow} (x)\
D_{p\uparrow - q\uparrow}(z)\
+\ q^{\downarrow} (x)\
D_{p\uparrow - q\downarrow}(z)\ \nonumber\\
&+& \bar{q}^{\uparrow} (x)\
D_{p\uparrow - \bar{q}\uparrow}(z)\
+\ \bar{q}^{\downarrow} (x)\
D_{p\uparrow - \bar{q}\downarrow}(z),
\label{norm}
\end{eqnarray}
where
\begin{eqnarray}
D(z)
&=& \sum_{s=-3/2}^{+3/2} D^{s}(z).
\end{eqnarray}
In relating the production rates for various polarized $\Delta^{++}$
we employ SU(6) spin-flavor wave functions, from which
simple relations among the valence diquark $\rightarrow$
$\Delta^{++}$ fragmentation functions, $D_{ qq_{j(j_z)} }^{s}(z)$,
can be deduced (the diquark state $qq_{j(j_{z})}$ is labeled by
its spin $j$ and spin projection $j_z$).
The leading functions are related by:
\begin{eqnarray}
D_{ uu_{1(1)} }^{\Uparrow}(z)
= 3\ D_{ uu_{1(1)} }^{\uparrow}(z)
= \frac{3}{2} D_{ uu_{1(0)} }^{\uparrow}(z)
= \frac{3}{2} D_{ uu_{1(0)} }^{\downarrow}(z), \label{leadFF}
\end{eqnarray}
with normalization determined from:
\begin{eqnarray}
D_{ uu_{1(1)} }^{\Uparrow}(z)
&=& \frac{3}{4} D_{ uu }(z). \label{polunp}
\end{eqnarray}
(Note that this is true only when the spin projections of the diquark
and $\Delta$ are aligned.)
The non-leading fragmentation functions are those which require
at least two $q \bar{q}$ pairs to be created from the vacuum, namely
$D_{ uu_{1(0)} }^{\Uparrow/\Downarrow}$,
$D_{ uu_{1(1)} }^{\downarrow/\Downarrow}$,
$D_{ ud_{0(0)} }^{\Uparrow/\uparrow/\downarrow/\Downarrow}$,
$D_{ ud_{1(0)} }^{\Uparrow/\uparrow/\downarrow/\Downarrow}$,
and
$D_{ ud_{1(1)} }^{\Uparrow/\uparrow/\downarrow}$,
and those which require 3 such pairs,
$D_{ uu_{1(1)} }^{\Downarrow}$ and
$D_{ ud_{1(1)} }^{\Downarrow}$.
Except at very small $z$ ($z \alt 0.2$) the latter functions
are consistent with zero \cite{BFM}.
For the 2-$q \bar{q}$ pair fragmentation functions, we also expect that
$ D_{ uu_{1(0)} }^{\Uparrow}(z)
= D_{ uu_{1(0)} }^{\Downarrow}(z)$.
For $z \agt 0.2$ the unpolarized model fragmentation functions
of Ref.\cite{BFM} requiring two $q \bar{q}$ pairs (e.g. $D_{ud}(z)$)
are quite small compared with the leading fragmentation functions,
$D_{ud}(z) \simeq 0.1\ D_{uu}(z)$.
For spin-dependent fragmentation we therefore expect
a similar behavior for those decay probabilities requiring two
$q \bar{q}$ pairs created in order to form the final
state with the correct spin and flavor quantum numbers.
This then allows for a complete model description of the
polarized fragmentation backgrounds
at large $z$ in terms of only the 4 fragmentation functions in
Eq.(\ref{leadFF}).
Finally, the $p_T$-integrated differential cross section
for the electroproduction of a $\Delta^{++}$ with spin $s$
can be written:
\begin{eqnarray}
{ d^3\sigma^{(s)} \over dx dQ^2 d\zeta }
&=&
\left( { 2 \pi \alpha^2 \over M^2 E^2 x (1-x) } \right)
\left( { 1 \over 2 x^2 }\
+\ { 4 M^2 E^2 \over Q^4 }
\left( 1 - {Q^2 \over 2 M E x} - { Q^2 \over 4 E^2 } \right)
\right)
\label{qpmful}\\
& & \hspace*{-1.3cm} \times
\left[
{4 x \over 9}
\left( u_V^{\uparrow} D_{ud_{1(0)}}^{s}
+ 2 \bar{u}^{\uparrow} \left({2 \over 3} D_{uu_{1(1)}}^{s}
+ {1 \over 3} D_{uu_{1(0)}}^{s}
\right)
+ u_V^{\downarrow} D_{ud_{1(1)}}^{s}
+ 2 \bar{u}^{\downarrow} \left({2 \over 3} D_{uu_{1(1)}}^{s}
+ {1 \over 3} D_{uu_{1(0)}}^{s}
\right)
\right)
\right.
\nonumber\\
& & \hspace*{-1.0cm} +
\left.
{x \over 9}
\left( d_V^{\uparrow} D_{uu_{1(0)}}^{s}
+ 2 \bar{d}^{\uparrow} \left({2 \over 3} D_{uu_{1(1)}}^{s}
+ {1 \over 3} D_{uu_{1(0)}}^{s}
\right)
+ d_V^{\downarrow} D_{uu_{1(1)}}^{s}
+ 2 \bar{d}^{\downarrow} \left({2 \over 3} D_{uu_{1(1)}}^{s}
+ {1 \over 3} D_{uu_{1(0)}}^{s}
\right)
\right)
\right].
\nonumber
\end{eqnarray}
In Figs.2a and 2b the parton model predictions (dashed) for the various
polarization states of the $\Delta$ are plotted in comparison
with the pion-exchange cross sections.
In the quark-parton model the correlations are significantly weaker,
with the ratio of polarized $\Delta$s being
$s = +3/2 : +1/2 : -1/2 : -3/2 \approx 3 : 2 : 1 : 0$.
The comparisons assume that there is no significant interference
between the parton fragmentation and pion-exchange contributions.
At small values of the exchanged four-momentum squared $t$ one may
expect this to be a good approximation, since the distance scales
at which the pion and diquark are formed are rather different.
For larger values of $t$ this approximation may be less justifiable,
and the possibility would exist that interference effects could
modify the above simple predictions.
This problem would be most pronounced for hard $\pi N \Delta$
vertices, however for relatively soft form factors
($\Lambda \alt 700$ MeV) the above predictions should be a reliable
guide.
Although the total unpolarized parton model cross sections
are larger than the pion-exchange cross sections, even at
larger values of $\zeta$ where the pionic effects are
strongest, the polarization aligned component in the pion
model is larger than that in the parton model.
The differences between the pion-exchange model and fragmentation
backgrounds can be further enhanced by examining polarization
asymmetries.
In Fig.4 we show the difference $\sigma^+ - \sigma^-$,
where
$\sigma^{\pm} \equiv Q^2 d^3\sigma^{(s=\pm 1/2)} / dx dQ^2 d\zeta $,
as a fraction of the total unpolarized cross section,
for the two kinematic cases in Figs.2a and 2b
(solid = CEBAF kinematics;
dashed = HERMES kinematics).
The resulting $\zeta$ distributions are almost flat, but
significantly different for the two models ($\pi$ and $qq$ label
the pion-exchange and spectator diquark fragmentation models).
We have also calculated the ratio for the form factor in
Eq.(\ref{FFt}), and find the results to be almost
indistinguishable from those in Fig.4.
Therefore, a measurement of the polarization asymmetry appears to
test only the presence of a pionic component of the nucleon
wave function, independent of the details of the form factor.
Of course the two curves in Fig.4 represent extreme cases,
in which $\Delta$s are produced entirely via pion emission or
diquark fragmentation.
In reality we can expect a ratio of polarization cross sections
which is some average of the curves in Fig.4.
The amount of deviation from the parton model curve will indicate
the extent to which the pion-exchange process contributes. From
this, one can in turn deduce the strength of the $\pi N \Delta$
form factor.
Unlike inclusive DIS, which can only be used to place upper bounds
on the pion number, the semi-inclusive measurements could pin down
the absolute value of $< n >_{\pi\Delta}$.
A measurement of this ratio would thus be particularly useful
in testing the relevance of non-perturbative degrees of freedom
in high energy processes.
\section{Kaon Cloud of the Nucleon}
Semi-inclusive leptoproduction of polarized $\Lambda$ hyperons from
polarized protons can also be used to test the relevance
of a kaon cloud in the nucleon, Fig.5.
The advantage of detecting $\Lambda$s in the final state, as compared
with $\Delta$ baryons lies in the fact that the $\Lambda$ is
self-analyzing.
It has, in fact, been suggested recently (see Ref.\cite{HMC}) that
measurement of the polarization of the $\Lambda$ in the target
fragmentation region could discriminate between models of the spin
content of the nucleon, in which a large fraction of the spin is
carried either by (negatively polarized) strange quarks or
(positively polarized) gluons.
The latter would imply a positive correlation of the
target proton and $\Lambda$ spins, while the spin projection
of the $\Lambda$ along the target polarization axis should be
negative in the former model.
(Similar effects would also be seen in the reaction
$\bar p p \rightarrow \bar \Lambda \Lambda$ \cite{AEK}.)
A kaon cloud would be the natural way to obtain a polarized strange
sea of the proton.
Although some data do exist for $\Lambda$ production in the region
$x_F < 0$ \cite{LAMDAT}, the large errors and limited range of $x_F$
do not permit one to unambiguously discern the presence of $K$ effects.
A direct test of the presence of a kaon cloud of the nucleon would be
to observe the differential $\Lambda$ production cross section
at large $\zeta$, and in particular the relative polarization yields.
The formalism for the DIS off the kaon cloud \cite{TTT} is very similar
to that for the pion exchange model in Section III.
Included in the observed $\Lambda$ cross section will be
contributions from direct $\Lambda$ production via $K^+$ exchange,
as well as those from $\Sigma^0$ recoil states, which subsequently
decay to $\Lambda \gamma$.
The differential hyperon $H$ ($= \Lambda, \Sigma^0$)
production cross section is similar to that in Eq.(\ref{ope5}), with
the trace factor here given by:
\begin{eqnarray}
\label{traceL}
{\cal T}_{H}^{S\ s}(t)
&=& \left( P \cdot p\ -\ M M_{H} \right)
\left( 1 + S \cdot s \right)\
-\ P \cdot s\ p \cdot S.
\end{eqnarray}
Using the rest frame parametrizations of the individual momentum
and spin vectors, we find for a proton initially polarized in the
positive $z$ direction:
\begin{eqnarray}
{\cal T}_{H}^{+{1\over 2}\ \pm{1\over 2}}(t)
&=& { 1 \over 2 } \left( (M_{H}-M)^2 - t \right)
\left( 1 \pm \cos\alpha \right).
\end{eqnarray}
The $\Lambda$ production cross section is shown in Fig.6,
as a function of $\zeta$, for the two possible polarizations
(the kinematics are as in Fig.2a).
The $K$-exchange predictions are calculated for the form factor
in Eq.(\ref{FF}) with cut-offs of $\Lambda = 0.6$ (smallest),
0.8 and 1.0 GeV (largest).
The $K$-exchange model predicts very strong correlations
between the target and recoil polarizations, so that the
asymmetry shown in Fig.7 for the cross sections
$(\sigma^+ - \sigma^-) / (\sigma^+ + \sigma^-)$,
where $\sigma^{\pm} \equiv Q^2 d^3\sigma^{(s=\pm 1/2)}/dx dQ^2 d\zeta$,
is almost unity.
The $K$-exchange ratios are very similar to the $\pi$-exchange
results in Fig.4, indicating the similar spin transfer dynamics
inherent in the meson cloud picture of the nucleon.
This is in strong contrast with the expectation from the
$qq \rightarrow \Lambda$ diquark fragmentation picture,
in which the target--recoil spin correlation is much weaker.
In fact, to a first approximation the $\Lambda^{\uparrow\downarrow}$
yields in the quark-parton model are equal.
Assuming an SU(6) symmetric wave function for the $\Lambda$,
namely $\Lambda^{\uparrow\downarrow}
\sim s^{\uparrow\downarrow} (ud)_{spin=0}$,
the leading fragmentation function will be
$D_{ud_{0(0)}}^{\Lambda^{\uparrow\downarrow}}$,
so that the relevant component of the SU(6) proton wave function
is $u^{\uparrow} (ud)_{spin=0}$.
Since one has equal probabilities to form a $\Lambda^{\uparrow}$
and $\Lambda^{\downarrow}$, in the leading fragmentation
approximation the asymmetry will be zero.
Of course, SU(6) symmetry breaking effects, as well as non-leading
fragmentation contributions, will modify this result, as will
contributions from the production and decay of
$\Sigma^{0 \uparrow\downarrow}$ hyperons (from the SU(6)
wave function one can see that a $\Sigma^{0 \uparrow}$ is more
likely to form from a $p^{\uparrow}$ than is a
$\Sigma^{0 \downarrow}$).
However, the qualitative result that the asymmetry is small should
remain true.
Therefore the observation of a large polarization asymmetry in the
large-$\zeta$ region of the target fragmentation region will be
evidence for a kaon-exchange fragmentation mechanism.
\section{OTHER TESTS OF PION EXCHANGE}
In addition to the above described process which may be tested
in upcoming experiments, several other novel ideas have been
proposed to identify a pionic component of the nucleon wave function.
In this section we will briefly outline a couple of them.
\subsection{Exclusive electroproduction of pions}
In a recent detailed analysis \cite{HEID}, it has been
suggested that measurements of fast pions in the final state
in coincidence with the final electron could be sensitive to
a pionic component of the nucleon.
Extending the exclusive analysis of G\"uttner et al. \cite{GCPP} in
the IMF, Pirner and Povh \cite{HEID} work within a constituent
quark picture in which the probability to find a pion in the
nucleon is expressed in terms of the pion distribution function
inside a constituent quark.
The differential pion-production cross section for the
``leading pion'' (integrated over transverse momenta)
is written as:
\begin{eqnarray}
{d\sigma \over dx dy dz}
&\propto& A(x,y,z)\ +\ B(x,y,z) F_{\pi}(Q^2)\
+\ C(x,y,z) F_{\pi}^2(Q^2),
\label{piprod}
\end{eqnarray}
where $z = E_{\pi}/\nu$ is the fraction of the photon's energy
carried by the pion,
and where the $A$ and $B$ terms describe soft and hard fragmentation,
respectively.
The function $C$ reflects coherent scattering from the
pion cloud of the constituent quark.
Each term in Eq.(\ref{piprod}) gives a characteristic $Q^2$-dependence,
namely $\log Q^2$, $1/Q^2$ and $1/Q^4$, respectively.
To isolate the coherent scattering from the pion one therefore
has to restrict oneself to the a region of not too high $Q^2$,
where the form factor suppression has not yet eliminated the
pion signal.
The useful observation in this analysis is that each of the
three processes has a quite distinct $z$-dependence.
The hard-fragmentation process gives a differential
cross section which is constant in $z$, and is important
in the intermediate $z$ region ($0.6 \alt z \alt 0.8$).
The soft fragmentation mechanism is dominant at small $z$,
$z \alt 0.6$, but dies out rather rapidly at larger $z$.
This fact may enable one to detect the pion-exchange process,
which dominates the region $0.8 \alt z \alt 1$, where it
predicts a contribution that is several times larger than
the constant-$z$, hard fragmentation mechanism.
The conclusion that the pion-exchange process is dominant is
consistent with our results above.
\subsection{Semi-inclusive scattering from $^3He$.}
Another novel idea was recently put forward by Dieperink and Pollock
\cite{DP}, where the suggestion was to measure the recoiling $^3He$
nucleus in deep-inelastic scattering from a $^3He$ target at
$x \sim 0.05 - 0.1$.
Unlike the exclusive experiments, one would not need to restrict
oneself to the small $Q^2$ region.
Because of the rather small probability for the nucleus to remain
intact after a hard interaction with a parton in one of the
constituent nucleons, backgrounds due to parton fragmentation
would be virtually eliminated.
Therefore the most likely mechanism responsible for the final state
$^3He$ nucleus would be DIS from a non-nucleonic component in the
target, the typical candidate being a pion.
Other non-nucleonic constituents could also give rise to
the same final state, such as DIS from a Pomeron in the diffractive
region.
However, in practice these could be eliminated by restricting oneself
to the specific kinematic region of not too small $x$.
An additional problem here in obtaining unambiguous information about
the $N \pi$ vertex from the $^3He\ \pi$ vertex would be nuclear effects
in the $^3He$ nucleus.
Furthermore, any final state interactions, leading to the
break up of the $^3He$ nucleus, could decrease the apparent number
of pions seen in the reaction, thus leading to an underestimate of
the pion multiplicity in the nucleon.
Nevertheless, this is an interesting idea, and a detailed study
should be performed with a view to determining the feasibility of
conducting this experiment in future.
\section{CONCLUSION}
We have outlined a series of semi-inclusive experiments on polarized
proton targets that may for the first time enable one to unambiguously
establish the presence of a pion and kaon cloud of the nucleon at
high energies.
The most difficult part of the calculation is the estimate of the
size of the competing diquark fragmentation process.
While we have used as much experimental data and theoretical
guidance as possible, in order to make that calculation reliable,
it could undoubtedly benefit from further study into the
polarized diquark $\rightarrow$ polarized baryon fragmentation
process.
Even bearing this caution in mind, our results (especially
Figs.4 and 7) are extremely encouraging.
While the experiments proposed here are difficult, requiring
all the intensity and duty factor one can obtain with modern
electron accelerators, it does seem that they will provide quite
clear information on the role of the pseudoscalar mesons in the
nucleon.
\acknowledgements
We would like to thank W. Weise for a careful reading of the manuscript.
This work was partially supported by the
Australian Research Council
and the BMFT.
W.M. would like to thank the University of Adelaide for its
hospitality during a recent visit, where this work was completed.
\references
\bibitem{GSR} Gottfried, K.:
Phys.Rev.Lett. {\bf 18}, 1174 (1967).
\bibitem{NMC} Amaudruz, P., et al. (New Muon Collaboration):
Phys.Rev.Lett. {\bf 66}, 2712 (1991);
Phys.Rev. D {\bf 50}, 1 (1994).
\bibitem{HM} Henley, E.M. and Miller, G.A.:
Phys.Lett. B {\bf 251}, 497 (1990).
\bibitem{SST} Signal, A.I., Schreiber, A.W. and Thomas, A.W.:
Mod.Phys.Lett. A {\bf 6}, 271 (1991).
\bibitem{MTS} Melnitchouk, W., Thomas, A.W. and Signal, A.I.:
Z.Phys. A {\bf 340}, 85 (1991).
\bibitem{KL} Kumano, S. and Londergan, J.T.:
Phys.Rev. D {\bf 44}, 717 (1991).
\bibitem{MA} Ma, B.-Q., Sch\"afer, A. and Greiner, W.:
Phys.Rev. D {\bf 47}, 51 (1993).
\bibitem{JUL} Hwang, W.-Y.P., Speth, J. and Brown, G.E.:
Z.Phys. A {\bf 339}, 383 (1991);
Szczurek, A. and Speth, J.:
Nucl.Phys. {\bf A555}, 249 (1993).
\bibitem{NA51} Baldit, A., et al. (NA51 Collaboration):
Phys.Lett. B {\bf 332}, 244 (1994).
\bibitem{EMCG1} Ashman, J., et al. (European Muon Collaboration):
Nucl.Phys. {\bf B328}, 1 (1989).
\bibitem{SMC} Adeva, B., et al. (Spin Muon Collaboration):
Phys. Lett. B {\bf 329}, 399 (1994).
\bibitem{SLACG1} Abe, K., et al. (E143 Collaboration):
Phys.Rev.Lett. {\bf 74}, 346 (1995).
\bibitem{U1GLU} Altarelli, G. and Ross, G.G.:
Phys.Lett. B {\bf 212}, 391 (1988);
Carlitz, R.D., Collins, J.C. and Mueller, A.H.:
Phys.Lett. B {\bf 214}, 229 (1988);
Efremov, A.V. and Teryaev, O.V.:
Dubna preprint JINR-E2-88-287 (1988);
Jaffe, R.L. and Manohar, A.:
Nucl.Phys. {\bf B337}, 509 (1990);
Steininger, K. and Weise, W.:
Phys.Rev. D {\bf 48}, 1433 (1993);
Bass, S.D. and Thomas, A.W.:
Prog.Part.Nucl.Phys. {\bf 33}, 449 (1994);
Narison, S., Shore, G.M. and Veneziano, G.:
Nucl.Phys. {\bf B433}, 209 (1995).
\bibitem{TTT} Signal, A.I. and Thomas, A.W.:
Phys.Lett. B {\bf 191}, 205 (1987).
\bibitem{EJ} Ellis, J. and Jaffe, R.L.:
Phys.Rev. D {\bf 9}, 1444 (1974).
\bibitem{SUL} Sullivan, J.D.:
Phys.Rev. D {\bf 5}, 1732 (1972).
\bibitem{T83} Thomas, A.W.:
Phys.Lett. {\bf 126} B, 97 (1983).
\bibitem{FMS} Frankfurt, L.L., Mankiewicz, L. and Strikman, M.I.:
Z.Phys. A {\bf 334}, 343 (1989).
\bibitem{HEID} Pirner, H.J. and Povh, B.:
in Proceedings of the Italian Physical Society,
Vol.44 (The ELFE project: an electron laboratory
for Europe),
ed. J. Arvieux and E. De Santis, 1992.
\bibitem{DP} Dieperink, A.E.L. and Pollock, S.J.:
Z.Phys. A {\bf 348}, 117 (1994).
\bibitem{STA} Stack, J.D.:
Phys.Rev.Lett. {\bf 28}, 57 (1972).
\bibitem{LUS} Lusignoli, M. and Srivastava, Y.:
Nucl.Phys. {\bf B138}, 151 (1978);
Lusignoli, M., Pistilli, P., and Rapuano, F.:
Nucl.Phys. {\bf B155}, 394 (1979).
\bibitem{CEBAF} Melnitchouk, W. and Thomas, A.W.:
in Proceedings of the Workshop on CEBAF at
Higher Energies,
eds. N.Isgur and P.Stoler (April 1994) p.359.
\bibitem{BEB77} Bebek, C.J., et al.:
Phys.Rev. D {\bf 15}, 3077 (1977).
\bibitem{ARN85} Arneodo, M., et al. (EM Collaboration):
Phys.Lett. B {\bf 150}, 458 (1985).
\bibitem{EMC86} Arneodo, M., et al. (EM Collaboration):
Nucl.Phys. {\bf B264}, 739 (1986).
\bibitem{E745} Kitagaki, T., et al. (E745 Collaboration):
Phys.Lett. B {\bf 214}, 281 (1988).
\bibitem{BEBC} Guy, J., et al. (BEBC Collaboration):
Phys.Lett. B {\bf 229}, 421 (1989).
\bibitem{IST} Ishii, C., Saito, K. and Takagi, F.:
Phys.Lett. B {\bf 216}, 409 (1989).
\bibitem{GRO} Bosveld, G.D., Dieperink, A.E.L. and Scholten, O.:
Phys.Lett. B {\bf 264}, 11 (1991);
Scholten, O. and Bosveld, G.D.:
Phys.Lett. B {\bf 265}, 35 (1991).
\bibitem{MTN} Melnitchouk, W., Thomas, A.W. and Nikolaev, N.N.:
Z.Phys. A {\bf 342}, 215 (1992).
\bibitem{MST} Melnitchouk, W., Schreiber, A.W. and Thomas, A.W.:
Phys.Rev. D {\bf 49}, 1183 (1994).
\bibitem{WEIN} Weinberg, S.:
Phys.Rev. {\bf 150}, 1313 (1966).
\bibitem{DLY} Drell, S.D., Levy D.J. and Yan, T.M.:
Phys.Rev. D {\bf 1}, 1035 (1970).
\bibitem{MTV} Melnitchouk, W. and Thomas, A.W.,
Phys.Rev. D {\bf 47}, 3794 (1993);
Thomas, A.W. and Melnitchouk, W.,
in: Proceedings of the JSPS-INS Spring School
(Shimoda, Japan),
(World Scientific, Singapore, 1993);
Melnitchouk, W.,
Ph.D. thesis, University of Adelaide, June 1993
(unpublished).
\bibitem{ZOL} Zoller, V.R.:
Z.Phys. C {\bf 54}, 425 (1992);
Holtmann, H., Szczurek, A. and Speth, J.:
J\"ulich preprint KFA-IKP(TH) 1993-33.
\bibitem{AG} Arakelyan, G. and A.Grigoryan, A.:
Sov.J.Nucl.Phys. {\bf 34}, 745 (1981).
\bibitem{MSM} Mulders, P.J., Schreiber, A.W. and Meyer, H.:
Nucl.Phys. {\bf A549}, 498 (1992).
\bibitem{GCPP} G\"uttner, F., Chanfray, G., Pirner, H.J. and Povh, P.:
Nucl.Phys. {\bf A429}, 389 (1984).
\bibitem{PISF} Shigetani, T., Suzuki, K. and Toki, H.:
Phys.Lett. B {\bf 308}, 383 (1993).
\bibitem{SHAKIN} Shakin, C.M. and Sun, W.-D.:
Phys.Rev. C {\bf 50}, 2553 (1994).
\bibitem{SMRS} Sutton, P.J., Martin, A.D., Roberts, R.G.
and Stirling, W.J.:
Phys.Rev. D {\bf 45}, 2349 (1992).
\bibitem{NA10} Betev, B., et al. (NA10 Collaboration):
Z.Phys. C {\bf 28}, 15 (1985).
\bibitem{SB} Brown, G.E., Buballa, M., Li, Z. and Wambach, J.:
Stony Brook preprint SUNY-NTG-94-54;
Buballa, M., preprint SUNY-NTG-94-61.
\bibitem{RS} Rarita, W. and Schwinger, J.:
Phys.Rev. {\bf 60}, 61 (1941).
\bibitem{BDM} Benmerrouche, M., Davidson, R.M. and Mukhopadhyay, N.C.:
Phys.Rev. C {\bf 39}, 2339 (1989).
\bibitem{FF} Field, R.D. and Feynman, R.P.:
Nucl.Phys. {\bf B136}, 1 (1978).
\bibitem{SSV} Sloan, T., Smadja, G. and Voss, R.:
Phys.Rep. {\bf 162}, 45 (1980).
\bibitem{SCHM} Schmitz, N.:
Int.J.Mod.Phys. A {\bf 3}, 1997 (1988).
\bibitem{GS} Gehrmann, T and Stirling, W.J.:
Z.Phys. C {\bf 65}, 461 (1995).
\bibitem{CTEQ} Lai, H.L., et al. (CTEQ):
preprint MSU-HEP-41024, hep-ph/9410404.
\bibitem{CK} Carlitz, R. and Kaur, J.;
Phys.Rev.Lett. {\bf 38} 674 (1977);
Kaur, J.:
Nucl.Phys. {\bf B128}, 219 (1977).
\bibitem{SCHAEFER} Sch\"afer, A.:
Phys.Lett. B {\bf 208}, 175 (1988).
\bibitem{RW} Renton, R. and Williams, W.S.C.:
Ann.Rev.Nucl.Part.Sci. {\bf 31}, 193 (1981).
\bibitem{BRE} Brenner, A.E., et al.:
Phys.Rev. D {\bf 26}, 1497 (1982).
\bibitem{ACK} Ackermann, H., et al.:
Nucl.Phys. {\bf B120}, 365 (1977).
\bibitem{SLO} Sukhatme, U.P., Lassila, K.E. and Orava, R.:
Phys.Rev. D {\bf 25}, 2975 (1982).
\bibitem{BFM} Bartl, A., Fraas, H., and Majoretto, W.:
Phys.Rev. D {\bf 26}, 1061 (1982).
\bibitem{FFFRAG} Field, R.D. and Feynman, R.P.:
Phys.Rev. D {\bf 15}, 2590 (1977).
\bibitem{LUND} Andersson, B., Gustafson, G., Ingelman, G.
and Sjostrand, T.:
Phys.Rep. {\bf 97}, 31 (1983).
\bibitem{FPS} Fontannaz, M., Pire, B. and Schiff, D.:
Phys.Lett. {\bf 77} B, 315 (1978);
Beavis, D. and Desai, B.R.:
Phys.Rev. D {\bf 23}, 1967 (1981).
\bibitem{COUNT} Brodsky, S.J. and Blankenbecler, R.:
Phys.Rev. D {\bf 10}, 2973 (1974);
Brodsky, S.J. and Farrar, G.:
Phys.Rev.Lett. {\bf 31}, 1193 (1975).
\bibitem{BFMS} Bartl, A., Fraas, H., and Majoretto, W.:
Z.Phys. C {\bf 6}, 335 (1980);
{\em ibid} {\bf 9}, 181 (1981).
\bibitem{BIGI} Bigi, I.I.Y.:
Nuov.Cim. {\bf 41A}, 43, 581 (1977).
\bibitem{DON} Donoghue, J.F.,
Phys.Rev. D {\bf 17}, 2922 (1978);
{\em ibid} D {\bf 19}, 2806 (1979).
\bibitem{HMC} Mallot, G., et al.:
Letter of intent, {\em Semi-inclusive muon scattering
from a polarized target},
preprint CERN/SPSLC 95-27 (March 1995).
\bibitem{AEK} Alberg, M., Ellis, J., and Kharzeev, D.:
Preprint CERN-TH/95-47 (February 1995).
\bibitem{LAMDAT} Arneodo, M., et al. (EM Collaboration):
Phys.Lett. {\bf 145} B, 156 (1984);
Hicks, R.G., et al.:
Phys.Rev.Lett. {\bf 45}, 765 (1980);
Cohen, I., et al.:
Phys.Rev.Lett. {\bf 40}, 1614 (1978);
Brock, R., et al.:
Phys.Rev. D {\bf 25}, 1753 (1982).
\newpage
\begin{figure}
\centering{\ \psfig{figure=zpa1.ps,height=11cm}}
\caption{Pion-exchange model of the semi-inclusive deep-inelastic
scattering from a polarized proton with a polarized recoil
$\Delta^{++}$ in the final state.}
\label{F1}
\end{figure}
\begin{figure}
\centering{\ \psfig{figure=zpa2a.ps,height=8cm}}
\centering{\ \psfig{figure=zpa2b.ps,height=8cm}}
\caption{Differential electroproduction cross section for various
polarization states of the $\Delta^{++}$, for typical
(a) CEBAF and (b) HERMES kinematics (see text).
The $\pi$-exchange model predictions (solid) are for cut-off
masses $\Lambda =$ 600 (smallest), 800 and 1000 (largest) MeV.
The top three solid curves are for spin $s=+1/2$ final states,
while the bottom three solid curves are for $s=-1/2$.
The quark-parton model background (dashed) is estimated using
the fragmentation functions extracted from the unpolarized
EMC data \protect\cite{EMC86} and Eqs.(\protect\ref{leadFF})
and (\protect\ref{polunp}).}
\label{F2}
\end{figure}
\begin{figure}
\centering{\ \psfig{figure=zpa3.ps,height=10cm}}
\caption{Background parton (spectator `diquark') fragmentation
process leading to the same $\Delta^{++}$ final state.}
\label{F3}
\end{figure}
\newpage
\begin{figure}
\centering{\ \psfig{figure=zpa4.ps,height=12cm}}
\caption{Polarization asymmetry for the $\pi$-exchange (upper curves)
and parton fragmentation (lower curves) models,
with $\sigma^{\pm}$ as defined in the text, and
$\sigma_{\rm tot}$ is the sum over all polarization states.
The solid and dashed lines are for CEBAF and HERMES
kinematics, respectively.}
\label{F4}
\end{figure}
\newpage
\begin{figure}
\centering{\ \psfig{figure=zpa5.ps,height=11cm}}
\caption{Kaon-exchange mechanism for the semi-inclusive production of
polarized $\Lambda$ hyperons.}
\label{F5}
\end{figure}
\newpage
\begin{figure}
\centering{\ \psfig{figure=zpa6.ps,height=8cm}}
\caption{Differential $\Lambda$ production cross section in the
$K$-exchange model, for form factor cut-offs
$\Lambda =$ 600 (smallest), 800 and 1000 (largest) MeV,
for $s=+1/2$ (upper three curves) and $s=-1/2$ (lower
three curves) final states.
Also included are contributions from $K^+ \Sigma^0$
states, with the subsequent decay
$\Sigma^0 \rightarrow \Lambda \gamma$.}
\label{F6}
\end{figure}
\begin{figure}
\centering{\ \psfig{figure=zpa7.ps,height=8cm}}
\caption{Polarization asymmetry for the $K$-exchange (solid)
model of $\Lambda$ production, compared with a leading
fragmentation approximation estimate for the parton
fragmentation process (dashed).}
\label{F7}
\end{figure}
\end{document}
|
\subsection*{Introduction}
It is a common practice in set theory that one is interested to
consider a generic extension $M_1$ of a model $M,$ after this a
generic extension $M_2$ of $M_1,$ and so on, including the case
of infinite or transfinite number of steps. Iterated forcing of
Solovay and Tennenbaum~\cite{st} allows us to engineer this
iterated construction in an ordinary one--step generic extension.
In the most of cases iterated forcing is used to define
transfinite sequences of models such that every model is a certain
generic extension of the preceding model. (We do not consider here
sophisticated details at limit steps). Identifying the steps of
this construction with ordinals, and interpreting the set of the
ordinals involved as the {\it support\/} or the {\it ``length''\/} of
the iteration, we may say that the classical iterated forcing is
an iterated forcing of {\it wellordered\/} ``length''.
In principle it does not require an essential improvement of the
basic iterater forcing method to define iterations of
{\it wellfounded\/}, but not linearly ordered, ``length''.
This version is much rarely used then
the basic one. (See Groszek and Jech~\cite{gj} for several
known applications.)
It is a much more challenging question (we refer to Groszek and
Jech~\cite{gj}, p.~6) to carry out ``ill''founded iterations. No
general method is known, at least.
For a few number of rather simple forcing notions, ``ill''founded
iterations can be obtained without any use of the idea of
iteration at all. For example if $a\in 2^\om$ is a Cohen generic
real over a model $M$ and $b=o(a)\in 2^m$ is defined for any
$a\in 2^\om$ by $b(m)=a(2m),\;\forall\,m,$ then the sequence
of reals $a_n$ defined by $a_0=a$ and $a_{n+1}=o(a_n)$ realizes
the iteration of Cohen forcing of ``length'' $\om^\ast$ (the reverse
order on natural numbers): every $a_n$ is Cohen generic over
$M[\ang{a_m:m>n}].$ This construction can be applied to Solovay
random reals as well.
An idea how to carry out iterated forcing of a linear but not
wellordered ``length'' $I$ can be as follows. let us first consider a
usual iteration of a ``length'' $\la\in\rbox{Ord}$ as a pattern to follow.
The forcing conditions in this case are functions $p$ defined on
the set $\la=\ans{\al:\al<\la}$ and satisfying certain property
$P(p,\al)$
for every $\al<\la.$ Now to proceed with the \dd\bI case one may
want to use functions $p$ defined on $I$ and satisfying $P(p,i)$
for all $i\in I$.
The principal problem in this argument is that in the wellordered
setting the property $P(p,\al)$ is itself defined by induction on
$\al$ in a quite sophisticated way. So we first have to eliminate
the induction and extend the property $P$ to ``ill''ordered sets.
We do not know how this can be realized at least for a more or
less representative category of forcing notions. There is,
however, a forcing which allows to express the property
$P$ in simple geometrical terms, so that the ``ill''founded
iterations become available. This is the perfect set forcing
introduced by
Sacks~\cite{sa}. (We refer to Baumgartner and Laver~\cite{bl} on
matters of iterated Sacks forcing, and Groszek~\cite{g} on
further applications.)
\pagebreak[3]
\bte
\label{m}
Let\/ ${\bbb{\bf M}\bbb}$ be a countable transitive model of $\ZFC,$ $\bI$ a
partially ordered\nopagebreak{}
set in\/ ${\bbb{\bf M}\bbb}.$ Then there exists a generic\/
\dd{\aleph_1}preserving extension\/
${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\bI}]$ of\/ ${\bbb{\bf M}\bbb}$ such that\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){{\rm\arabic{enumi}}}
\item\label{m1}
For every\/ $\bald{i}\in \bI,$ ${\bbsp\bbox{a}\bbsp}_\bald{i}$ is a Sacks--generic real
over\/ ${\bbb{\bf M}\bbb}[\ang{{\bbsp\bbox{a}\bbsp}_\bald{j}:\bald{j}<\bald{i}}]$.\vspace{-1mm}
\item\label{m1+}
If\/ $\bald{i},\,\bald{j}\in\bI$ and\/ $\bald{i}<\bald{j}$ then\/ ${\bbsp\bbox{a}\bbsp}_\bald{i}\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{a}\bbsp}_\bald{j}]$.\vspace{-1mm}
\item\label{m2}
If\/ $\xi\in{\bbb{\bf M}\bbb}$ is an initial segment in\/ $\bI$ and\/
$\bald{i}\in\bI\setminus\xi$ then\/
${\bbsp\bbox{a}\bbsp}_\bald{i}\not\in{\bbb{\bf M}\bbb}[\ang{{\bbsp\bbox{a}\bbsp}_\bald{j}:\bald{j}\in\xi}]$.\vspace{-1mm}
\item\label{m2+}
If\/ $\xi\in{\bbb{\bf M}\bbb}$ is a countable in\/ ${\bbb{\bf M}\bbb}$ initial segment in\/
$\bI$ and\/
$c$ is a real in\/ ${\bbb{\bf N}\bbb}$ such that\/ ${\bbsp\bbox{a}\bbsp}_\bald{i}\in{\bbb{\bf M}\bbb}[c]$ for all\/
$\bald{i}\in\xi$ then the indexed set\/ $\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\xi}$
belongs to\/ ${\bbb{\bf M}\bbb}[c]$.\vspace{-1mm}
\item\label{m3}
For any initial segment\/ $\xi\subseteq\bI,$ $\xi\in{\bbb{\bf M}\bbb},$ and any real\/
$c\in{\bbb{\bf N}\bbb},$ we have exactly one from the following:\hspace{4mm}
\begin{minipage}[t]{0.6\textwidth}
{\rm(a)} $c\in {\bbb{\bf M}\bbb}[\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\xi}]$, \hspace{10mm}
or\vspace{2mm}
{\rm(b)} \hspace{1mm}there exists\/ $\bald{i}\in\bI\setminus\xi$
such that ${\bbsp\bbox{a}\bbsp}_\bald{i}\in {\bbb{\bf M}\bbb}[c]$.
\end{minipage}
\hfill $\,$
\end{enumerate}
\ete
The set $\bI$ is not necessarily wellfounded or linearly ordered
in ${\bbb{\bf M}\bbb}.$ In the particular case of inverse ordinals taken as
$\bI$ the theorem was recently proved by Groszek~\cite{g94}.
Items \ref{m2}, \ref{m2+}, \ref{m3} seem to show that the
\dd{\bbb{\bf M}\bbb} degree of a real $c$ in the extension ${\bbb{\bf N}\bbb}$ intends to be
determined by the set $\bI_c=\ans{\bald{i}\in\bI:{\bbsp\bbox{a}\bbsp}_\bald{i}\in{\bbb{\bf M}\bbb}[c]},$ an
initial segment of $\bI$ by item~\ref{m1+}. One can easily prove
that in fact $\bI_c=\bI_{c'}$ implies ${\bbb{\bf M}\bbb}[c]={\bbb{\bf M}\bbb}[c']$ provided
the set $\bI_c=\bI_{c'}$ belong to ${\bbb{\bf M}\bbb}$ (e.g. in the case when
all initial segments of $\bI$ belongs to ${\bbb{\bf M}\bbb}$), but the general
case remains open.
The proof of the theorem is based on a version of iterated Sacks
forcing realized in the form of perfect sets~\footnote
{\rm\ We consider perfect sets rather than perfect trees, because
the particular combinatorial properties we need hardly can be
expressed in a reasonable form for the forcing realized using
trees rather than perfect sets. Of course the absoluteness of the
conditions is lost because a perfect set in ${\bbb{\bf M}\bbb}$ is not perfect
both in the universe and the extension, but this is a comparably
minor problem, easily fixed by taking the topological closure.}
with certain combinatorial properties.
We shall be mostly concentrated on the case when $\bI$ is
{\it finite or countable\/} in ${\bbb{\bf M}\bbb}.$ The forcing we use in this
case will be a collection $\perfm$ of perfect subsets of the
product in ${\bbb{\bf M}\bbb}$ of $\bI$ copies of the Cantor space $2^\om,$
\dd\bald{i} th copy being
responsible for the corresponding ${\bbsp\bbox{a}\bbsp}_\bald{i}.$ Sections \ref{prelim}
through \ref{cont} of the paper present useful properties of sets
in $\perfm$ and continuous functions defined on them. This part of
the paper is not related to any particular model but
finally the reasoning will be ``relativized'' to ${\bbb{\bf M}\bbb}$.
The results of this study are used in sections \ref{re} through
\ref{uncount} for the proof of Theorem~\ref{m}. We show that the
reals in a \dd\mathord{{\rm I}\hspace{-2.5pt}{\rm P}} generic extension can be presented by continuous
functions in the ground model ${\bbb{\bf M}\bbb},$ defined on sets $X\in\perfm.$
It occurs that notions related to degrees of \dd{\bbb{\bf M}\bbb}
constructibility of reals in the extension are adequately
reflected in properties of continuous functions in the ground
model.
The case of {\it arbitrary\/} $\bI$ is reduced in
Section~\ref{uncount} to the case of countable $\bI$ by an
ordinary ``countable support'' argument.
\subsubsection*{An application: non--Glimm--Effros $\is12$
equivalence}
Harrington, Kechris, and Louveau~\cite{hkl} proved that each
Borel equivalence relation $\mathbin{\relf{E}}$ on reals satisfies one and
only one of the following conditions:\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){{\rm\hskip2pt(\Roman{enumi})\hskip2pt}}
\def\theenumi{\arabic{enumi})}
\item\label{1} \hspace{-1\mathsurround}
$\mathbin{\relf{E}}$ admits a countable Borel separating family.\vspace{-1mm}
\item\label{2} \hspace{-1\mathsurround}
$\mathbin{\relf{E}}$ continuously embeds $\mathbin{\relf{E}_0},$ the Vitali equivalence.
\end{enumerate}
(Some notation. A {\it separating family\/} for an equivalence
$\mathbin{\relf{E}}$ on reals is an indexed family $\ang{X_\al:\al<\gamma}$
($\gamma\in\rbox{Ord}$)
of sets $X_\al$ such that $x\mathbin{\relf{E}} y$ iff
$\forall\,\al\,(x\in X_\al\,\llra\,y\in X_\al)$ for all $x,\,y.$
$\mathbin{\relf{E}_0}$ is the {\it Vitali equivalence\/} on the
{\it Cantor space\/} ${\skri D}=2^\om,$ defined
by: ${x\mathbin{\relf{E}_0} y}$ iff $x(n)=y(n)$ for all but finite $n\in\om.$ An
{\it embedding\/} of $\mathbin{\relf{E}_0}$ into $\mathbin{\relf{E}}$ is a $1-1$ function
$U:{\skri D}\,\lra\,\hbox{reals}$ such that
$x\mathbin{\relf{E}_0} y\;\llra\;U(x)\mathbin{\relf{E}} U(y)$ for all $x,\,y\in {\skri D}$.
We refer the reader to \cite{hkl} as the basic sourse of
information on the matter.)
Hjorth and Kechris~\cite{hk}, Hjorth~\cite{h-det},
Kanovei~\cite{k-sm,k-s11} obtained partial results of
this type for $\fs11$ and even more complicated relations, which
we do not intend to discuss here.
However there exists a $\is12$ equivalence relation which does
not admit a theorem of the Glimm--Effros type in $\ZFC,$ at least
in the field of real--ordinal definable (R-OD, in brief)
separating families and embeddings.
\bte
\label{ge}
It is consistent with\/ $\ZFC$ that the\/ $\is12$ equivalence
relation\/ $\mathbin{\relf{C}}$ defined on reals by\/ $x\mathbin{\relf{C}} y$ iff\/
$\rbox{L}[x]=\rbox{L}[y]\;:$\vspace{-1mm}
\begin{itemize}
\item[--] neither has a R-OD separating family$;$
\vspace{-1mm}
\item[--] nor admits an uncountable R-OD pairwise\/
\dd\mathbin{\relf{C}} inequivalent set$.$
\end{itemize}
\ete
\noindent
{\bf Remarks} \
1. The ``nor'' part of the theorem implies that
$\mathbin{\relf{C}}$ does not embed $\mathbin{\relf{E}_0}$ via a \hbox{R-OD} embedding, because
obviously there exists a perfect set of pairwise \dd\mathbin{\relf{E}_0}
inequivalent points.
2. It makes no sense to look for non-R-OD separating families
in the ``either'' part. Indeed let $\kappa$ be
the cardinal of the quotient set $\hbox{reals}/\mathbin{\relf{C}}.$ Then any
enumeration $\ang{X_\al:\al<\kappa}$ of all \dd\mathbin{\relf{E}} equivalence
classes is a separating family, but this construction does not
guarantee the real--ordinal definability of the enumeration even
in the case when $\mathbin{\relf{E}}$ itself is R-OD (take $\mathbin{\relf{E}_0}$ as an example).
\vspace{4mm}
The model for Theorem~\ref{ge} we propose is the iterated Sacks
extension of the constructible model having $\om_1\times\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}$
($\om_1$ copies of the integers) as the ``length'' of iteration.
The model is considered in Section~\ref{seqge}.
\newpage
\subsection{Notation and pre--conditions}
\label{prelim}
\label{prop}
\label{splt}
{\it The ``length''\/}. Let $\bI$ be a fixed countable partially
ordered set, which will be the ``length'' of iteration. Characters
$\bald{i},\,\bald{j}$ are used to denote elements of $\bI.$ Subsets of $\bI$
will be denoted by Greek letters $\xi,\,\eta,\,\zeta$.\vspace{1mm}\vom
{\it Spaces\/}. ${\skri N}=\om^\om$ is the {\it Baire space\/}; points of
${\skri N}$ will be called {\it reals\/}. ${\skri D}=2^\om$ is the
{\it Cantor space\/}. For
$\xi\subseteq \bI,$ $\can\xi$ is the product of \dd\bI many copies of
${\skri D}$ with the product topology (here $\bI$ is considered as
discrete). Then every $\can\xi$ is a compact space homeomorphic
to ${\skri D}$ itself unless $\xi=\emptyset$.\vspace{1mm}\vom
{\it Projections\/}. Assume that $\xi\subseteq\eta\subseteq\bI.$ If
$x\in\can\eta$ then let $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\in\can\xi$ denote the
usual restriction. If $X\subseteq\can\eta$ then let
$X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi=\ans{x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi:x\in X}$.
But if $X\subseteq\can\xi$ then we set
$X\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt} \eta=\ans{y\in\can\eta:y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi\in X}$.
In addition, if $\bald{i}\in\xi\subseteq\bI$ and $X\subseteq\can\xi$ then we put
$X(\bald{i})=\ans{x(\bald{i}):x\in X}$.\vspace{1mm}\vom
{\it Initial segments\/}. Let $\IS$ denote the set of all
initial segments of $\bI.$
For any $\bald{i}\in\bI,$ we put $[<\hspace{-2pt}\bald{i}]=\ans{\bald{j}\in\bI:\bald{j}<\bald{i}},$
$[\not>\hspace{-2pt}\bald{i}]=\ans{\bald{j}\in\bI:\bald{j}\not>\bald{i}},$ and $[\<\hspace{-2pt}\bald{i}],$
$[\not\>\hspace{-2pt}\bald{i}]$ in the same way.
To save space, let $X\rsd{<\bald{i}}$ mean $X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}[<\hspace{-2pt}\bald{i}],$
$\can{\<\bald{i}}$ mean $\can{[\<\bald{i}]},$ etc.\vspace{1mm}\vom
If $\bald{i}\in\xi\in\IS$ and $X\subseteq\can\xi,$ then we
define $D_{Xz}(\bald{i})=\ans{x(\bald{i}):x\in X\;\,\&\;\, z=x\rsd{<\bald{i}}}$
for every $z\in X\rsd{<\bald{i}}.$ Thus $D_{Xz}(\bald{i})\subseteq {\skri D}$.\vspace{1mm}\vom
{\it Pre--conditions\/}.
The following definition would be sufficient for the purpose to
prove Theorem~\ref{m} at least in two particular cases: when $\bI$
is wellfounded, and when $\bI$ is linearly ordered. In fact we
don't know whether it gives the expected result in general case.
We are not able to prove a very important technical fact
(Proposition~\ref{clop} below): if $X'$ is a clopen (in the
relative topology) nonempty subset of $X\in\ipe\xi$ then $X'$
contains a subset $X''\in\ipe\xi.$ This is why one more
requirement will be added in Section~\ref{forc}, to define the
notion of forcing completely.
\bdf
(Pre--conditions)\\[1pt]
For any $\zeta\in\IS,$ $\ipe\zeta$ is the collection of
all sets $X\subseteq\can\zeta$ such that\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){{\rm P-\arabic{enumi}}}
\item The set $X$ is closed and nonempty.\vspace{-1mm}
\item\label{perf1}
If $\bald{i}\in\zeta$ and $z\in X\rsd{<\bald{i}}$ then
$D_{Xz}(\bald{i})$ is a perfect set in ${\skri D}$.\vspace{-1mm}
\item\label{oz}
If $\bald{i}\in \zeta$ and $G\subseteq{\skri D}$ is open then the set
$\ans{x\rsd{<\bald{i}}:x\in X\;\,\&\;\, x(\bald{i})\in G}$ is open in
$X\rsd{<\bald{i}}$.~\footnote
{\rm\ In other words, it is required that the projection from
$X\rsd{\<\bald{i}}$ to $X\rsd{<\bald{i}}$ is an open map.}\vspace{-1mm}
\item\label{indep}
If $\xi,\,\eta\in\IS,$
$\xi\cup \eta\subseteq \zeta,$ $x\in X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi,$ $y\in X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \eta,$
and $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cap \eta)=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cap \eta),$ then
$x\cup y\in X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cup \eta)$.\vspace{-1mm}
\end{enumerate}
Finally we set $\perf'=\ipe\bI$.\qed
\edf
This section contains several quite elementary lemmas on
pre--conditions.
\bass
\label{less'}
If\/ $X\in\ipe\zeta$ and\/ $\xi\in\IS,$ $\xi\subseteq \zeta,$ then\/
$X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi\in\ipe\xi$.\qed
\eass
\ble
\label{pro'}
Suppose that\/ $\xi,\,\eta,\,\zeta\in\IS,$ $\xi\cup \eta\subseteq \zeta,$
$X\in\ipe\zeta,$ $Y\subseteq X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \eta,$ and\/ $Z=X\cap (Y\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\zeta).$
Then $Z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi=(X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi)\cap (Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cap \eta)\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt} \xi)$.
\ele
\noindent{\bft Proof\hspace{2mm} } The inclusion $\subseteq$ is quite easy. To prove the opposite
direction let $x$ belong to the right--hand side. Then in
particular $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cap \eta)=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cap \eta)$ for some
$y\in Y.$ On the other hand $x\in X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and $y\in X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta.$
Condition~\ref{indep} implies $x\cup y\in X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cup \eta).$
Therefore $x\cup y\in Z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cup \eta)$ because $y\in Y.$ We
conclude that $x\in Z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi$.\qed
\ble
\label{apro'}
Suppose that\/ $\xi,\,\zeta\in\IS,$ $\xi\subseteq \zeta,$ $X\in\ipe\zeta,$
$Y\in\ipe\xi,$ and\/ $Y\subseteq X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi.$ Then\/
$Z=X\cap (Y\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt} \zeta)$ belongs to $\ipe\zeta$.
\ele
\noindent{\bft Proof\hspace{2mm} } We check condition~\ref{perf1}. Let $\bald{i}\in\zeta$ and
$z\in Z\rsd{<\bald{i}}.$ If $\bald{i}\in \xi$ then obviously
$D_{Zz}(\bald{i})=D_{Yz}(\bald{i}).$ If $\bald{i}\in \zeta\setminus\xi$ then
$D_{Zz}(\bald{i})=D_{Xz}(\bald{i})$ by Lemma~\ref{pro'} (for
$\eta=[\<\hspace{-2pt}\bald{i}]$).\vspace{1mm}
We check \ref{oz}. Let $\bald{i}\in\zeta.$ The case $\bald{i}\in\xi$ is easy as
above, so let us suppose that $\bald{i}\in \zeta\setminus \xi.$ We assert
that\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){(\arabic{enumi})}
\def\theenumi{\arabic{enumi})}
\item\label{aa1} \mathsurround=1mm
$Z\rsd{<\bald{i}}=(X\rsd{<\bald{i}})\cap (Y'\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}[<\hspace{-2pt}\bald{i}]),$ where
$Y'=Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi'$ and $\xi'=\xi\cap[<\hspace{-2pt}\bald{i}]$, \hfill and\vspace{-1mm}
\item\label{aa2} \mathsurround=1mm
$\ans{z\in Z:z(\bald{i})\in G}\rsd{<\bald{i}}=
(\ans{x\in X:x(\bald{i})\in G}\rsd{<\bald{i}})\cap (Y'\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt} [<\hspace{-2pt}\bald{i}])$.\vspace{-1mm}
\end{enumerate}
Indeed \ref{aa1} immediately follows from Lemma~\ref{pro'}.
The direction $\subseteq$ in \ref{aa2} is obvious. To prove the
opposite direction, let $z\in\can{<\bald{i}}$ belong to the right--hand
side, so that $z=x\rsd{<\bald{i}}$ for some $x\in X$ such that
$x(\bald{i})\in G,$ and $z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'$ for some $y\in Y.$
Applying property \ref{indep} of $X,$ we get $x'\in X$ such that
$x'\rsd{\<\bald{i}}=x\rsd{\<\bald{i}}$ and $x'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=y.$ In particular, we
have $x'\in Z$ and $x'(\bald{i})\in G,$ so that
$z=x\rsd{<\bald{i}}=x'\rsd{<\bald{i}}$ belongs to the left--hand side.
Thus both \ref{aa1} and \ref{aa2} are verified. Now it
suffices to recall that $X$ satisfies~\ref{oz}, which implies that
the set $\ans{x\in X:x(\bald{i})\in G}\rsd{<\bald{i}}$ is clopen in
$X\rsd{<\bald{i}}$.\vspace{1mm}
We check \ref{indep}. Assume that $\eta,\,\tau\in\IS,$
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta\in Z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,$ and $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\tau\in Z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\tau;$ we have
to prove that $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta\cup\tau)\in Z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta\cup\tau).$ Let
$\eta'=\xi\cap\eta$ and $\tau'=\xi\cup\tau$.
First of all we note that
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta\cup\tau)\in X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta\cup\tau)$ by
property~\ref{indep} of $X.$ Then, since it follows from
Lemma~\ref{pro'} that
${Z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta\cup\tau)=(X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta\cup\tau))\,\cap\,
(Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta'\cup\tau')\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}(\eta\cup\tau))},$
it suffices to verify that
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta'\cup\tau')\in Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\eta'\cup\tau').$
But $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta'\in Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta'$ by the choice of $x,$ and
the same for $\tau',$ so that the required fact follows from
property~\ref{indep} of $Y$.\qed
\ble
\label{gath'}
Let\/ $\bald{i}\in\bI,$ $X\in\ipe{\<\bald{i}},$ $X'\subseteq X$ is closed and
nonempty, $X'\rsd{<\bald{i}}=X\rsd{<\bald{i}},$ and\/ $X'$ satisfies\/
{\rm\ref{perf1}} and\/ {\rm\ref{oz}} for this
particular\/ $\bald{i}.$ Then\/ $X'\in\ipe{\<\bald{i}}$.
\ele
\noindent{\bft Proof\hspace{2mm} } Notice that requirement \ref{indep} is automatically
satisfied in this case provided either $\xi$ or $\eta$ contains
$\bald{i}$.\qed
\subsubsection*{Splitting of pre--conditions}
We now demonstrate how a set in $\ipe\zeta$ can be
splitted in a pair of smaller sets. Let $A\subseteq{\skri D}$ be a set
containing at least two different points. The largest finite
sequence $r\in 2^{<\om}$ such that $r\subset a$ for all $a\in A$
is denoted by $\rbox{root}\hspace{1pt}(A).$ We put $\rbox{stem}\hspace{1pt}(A)=\rbox{dom}\,\rbox{root}\hspace{1pt}(A)$ and define
$$
\rbox{Spl}(A,e)=\ans{a\in A:a(l)=e},\hspace{5mm}\hbox{where}
\hspace{3mm}l=\rbox{stem}\hspace{1pt}(A)\hspace{3mm}\hbox{and}\hspace{3mm}
e=0\hspace{2mm}\hbox{or}\hspace{2mm}1\,.~\footnote
{\rm\ Digits $0$ and $1$ will be denoted usually by letters
$e$ and $d$.}
$$
Let now $X\in\ipe\zeta.$ Suppose that $\bald{i}\in\zeta.$ For any
$y\in Y=X\rsd{<\bald{i}}$ the set $A(y)=D_{Xy}(\bald{i})\subseteq{\skri D}$ is perfect;
therefore so are the sets $\rbox{Spl}(A(y),e),\,\,e=0,\,1.$ We define
$$
\rbox{Spl}(X,\bald{i},e)=\ans{x\in X:x(\bald{i})\in \rbox{Spl}(A(x\rsd{<\bald{i}}),e)}
\hspace{5mm}\hbox{for}\hspace{3mm}e=0,\,1\,.
$$
\ble
\label{spl'}
Assume that\/ $\zeta\in\bI,$ $\bald{i}\in\zeta,$ and\/ $X\in\ipe\zeta$.
Then\vspace{-1mm}
\begin{enumerate}
\item
The sets $X_e=\rbox{Spl}(X,\bald{i},e),\;e=0,\,1,$ belong to $\ipe\zeta$.\vspace{-1mm}
\item
$X_0{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\zeta\cap[\not\>\hspace{-2pt}\bald{i}])=X_1{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\zeta\cap[\not\>\hspace{-2pt}\bald{i}])=
X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\zeta\cap[\not\>\hspace{-2pt}\bald{i}])$.\vspace{-1mm}
\item $X_0\rsd{\<\bald{i}}\cap X_1\rsd{\<\bald{i}}=\emptyset$.
\end{enumerate}
\ele
\noindent{\bft Proof\hspace{2mm} } First of all we note that since
$X_e=X\cap(\rbox{Spl}(X\rsd{\<\bald{i}},\bald{i},e)\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\zeta),$ Assertion~\ref{less'}
and Lemma~\ref{apro'} allow to consider only the case
$\zeta=[\<\hspace{-2pt}\bald{i}].$ In this case items 2 and 3 become obvious, so
we concentrate on item 1.
To prove that $X_e\in\ipe{\<\bald{i}},$ we put $Y=X\rsd{<\bald{i}},$
$D(y)=D_{Xy}(\bald{i}),$ and then define
$Y_r=\ans{y\in Y:\rbox{root}\hspace{1pt}(D(y))=r}$ for all
$y\in Y$ and $r\in 2^{<\om}.$ It is implied by property {\ref{oz}}
of $X$ that the sets $Y_r$ are clopen in $Y,$ and in fact there
exist only finitely many nonempty sets $Y_r.$ Therefore the sets
$$
X_e={\textstyle\bigcup_r}\ans{x\in X:x\rsd{<\bald{i}}\in Y_r\;\,\&\;\,
x(\rbox{dom}\, r)=e}\,,\hspace{5mm}e=0,\,1\,,
$$
are clopen in $X,$ and $X_e\rsd{<\bald{i}}=X\rsd{<\bald{i}}.$
Furhermore condition \ref{perf1} for $X_e$ for the given $\bald{i}$
follows from the fact that nonempty intersections of perfect and
clopen sets are perfect. Finally condition \ref{oz} for $X_e$ for
the given $\bald{i}$ can be easily obtained from \ref{oz} for $X$ using
the decomposition given by the last displayed
formula.\qed
\newpage
\subsection{The forcing}
\label{forc}
The splitting procedure plays principal role in the complete
definition of the notion of forcing. Let us start with several
auxiliary definitions.
\bdf
Let $\zeta\subseteq\bI.$ A \dd\zeta{\it admissible\/} function is a function
$\Phi:\om\;\lra\;\zeta$ taking each value $\bald{i}\in\zeta$ infinitely many
times.\qed
\edf
\bdf
Assume that $\zeta\in\IS$ and $\Phi$ is a \dd\zeta admissible
function. Let $X\in\ipe\zeta.$ We define a set $X[u]=X_\Phi[u]$ for
all $u\in 2^{<\om}$ as follows.\vspace{-1mm}
\begin{enumerate}
\item $X[\La]=X$. ($\La$ is the empty sequence, the only member
of $2^0$.)\vspace{-1mm}
\item If $m\in\om$ $u\in 2^m,$ and $X[u]$ has been defined, we
put $X[u{\mathbin{\kern 1.3pt ^\wedge}} e]=\rbox{Spl}(X[u],\bald{i},e),$ where $\bald{i}=\Phi(m),$ for
$e=1,\,2$.\vspace{-1mm}
\end{enumerate}
For an infinite sequence $a\in 2^\om$ we define
$X[a]=X_\Phi[a]=\bigcap_{m\in\om} X[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m]$. (Notice that
$X[a]$ is nonempty by the compactness of $\can\zeta$.)\qed
\edf
\begin{corollary}\TF\
\label{itspl'}
If\/ $X\in\ipe\zeta$ then\/
$X_\Phi[u]\in\ipe\zeta$ for all\/ $u\in 2^{<\om}$.\qed
\end{corollary}
\bdf
(Forcing conditions)\\[1pt]
Let $\zeta\in\IS.$ A set $X\in\ipe\zeta$ is {\it shrinkable\/} if for
any \dd\zeta admissible function $\Phi$ and any $a\in 2^\om,$ the
set $X_\Phi[a]$ contains only one point.
We put $\pe\zeta=\ans{X\in\ipe\zeta:X\,\,\hbox{is shrinkable}}$ and
$\perf=\pe\bI$.\qed
\edf
\noindent
Obviously $X=\can\zeta$ is shrinkable (and belongs to $\perf$).
We can easily prove that if $\zeta$ is wellfounded then every
$X\in\ipe\zeta$ is shrinkable, so that $\pe\zeta=\ipe\zeta.$ On the
other hand if $\zeta=\om^\ast$ (the order of negative integers)
then the set
$$
X=\ans{x\in\can\zeta:\forall\,\bald{i}\in\zeta\,(x(\bald{i})(0)=0)\,\hbox{ or }\,
\forall\,\bald{i}\in\zeta\,(x(\bald{i})(0)=1)}
$$
is not shrinkable: every set $X_\Phi[a]$ contains two different
points.
We now show that the lemmas already proved for pre--conditions
remain valid for sets in $\pe\zeta$ --- {\it conditions\/} ---
as well.
\bass
\label{less}
If\/ $X\in\pe\zeta$ and\/ $\xi\in\IS,$ $\xi\subseteq \zeta,$ then\/
$Y=X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi\in\pe\xi$.
\eass
\noindent{\bft Proof\hspace{2mm} } We have only to verify that $Y$ is shrinkable. Let $\Psi$
be a \dd\xi admissible function, and $b\in 2^\om.$ We want to
prove that $Y_\Psi[b]$ is a singleton. Let $\Phi$ be a \dd\zeta
admissible function such that $\Phi(2n)=\Psi(n)$ and
$\Phi(2n+1)\in\zeta\setminus\xi$ for all $n.$ Let $a\in 2^\om$ be
defined by $a(2n)=b(n)$ and $a(2n+1)=0$ for all $n.$ Then
$X_\Phi[a]$ is a singleton since $X$ is shrinkable; on the other
hand, $Y_\Psi[b]=X_\Phi[a]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$.\qed
\ble
\label{apro}
Suppose that\/ $\xi,\,\zeta\in\IS,$ $\xi\subseteq \zeta,$ $X\in\pe\zeta,$
$Y\in\pe\xi,$ and\/ $Y\subseteq X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi.$ Then\/ $Z=X\cap(Y\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\zeta)$
belongs to $\pe\zeta$.
\ele
\noindent{\bft Proof\hspace{2mm} } To verify shrinkability let $\Phi$ be a \dd\zeta admissible
function and ${a\in 2^\om}.$ To prove that $Z_\Phi[a]$ is a
singleton, let ${\Om=\ans{k:\Phi(k)\in\xi}=\ans{o_m:m\in\om}}$ in
the increasing order. We define $\Psi(m)=\Phi(o_m),$ so that
$\Psi$ is \dd\xi admissible. We also put $b(m)=a(o_m).$ Then
$Y_\Psi[b]$ is a singleton; on the other hand, $Y_\Psi[b]=
Z_\Phi[a]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ because splittings $\rbox{Spl}(...,\bald{i},e)$ with
$\bald{i}\in\zeta\setminus\xi$ do not change projections on $\xi$.
Thus at least $Z_\Phi[a]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=\ans{y}$ is a singleton.
Let now $Y'=X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ so that $Y\subseteq Y'$ and $y\in Y'.$ By the
shrinkability of $X,$ there exists some $a'\in 2^\om$ such that
$y=x'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ where $x'$ is the only element of $X_\Phi[a'].$
We now compose $c\in 2^\om$ from $a'$ and $a$ as follows:
$c(k)=a'(k)$ for $k\in\Om$ and $c(k)=a(k)$ for
$k\in\om\setminus\Om$.
Since $c{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\Om=a'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\Om,$ we can easily prove that
$X_\Phi[c]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X_\Phi[a']{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=\ans{y}=Z_\Phi[a].$
Furthermore since $c{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\om\setminus\Om)=a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\om\setminus\Om),$
we obtain that in general $X_\Phi[c]=Z_\Phi[a],$ as required.\qed
\ble
\label{gath}
If\/ $X\in\pe{\<\bald{i}}$ in
Lemma~\ref{gath'} then\/ $X'$ also
belongs to $\pe{\<\bald{i}}$.
\ele
\noindent{\bft Proof\hspace{2mm} } $X'\rsd{<\bald{i}}=Y=X\rsd{<\bald{i}}$ is shrinkable by
Lemma~\ref{less}. On the other hand, for any $y\in Y$ the set
$D_{X'y}(\bald{i})=\ans{x(\bald{i}):x\in X'}$ is converted to a singleton
after infinitely many operations $\rbox{Spl}(...,\bald{i},e)$\qed
\ble
\label{spl}
Assume that\/ $X\in\pe\zeta$ in Lemma~\ref{spl'}. Then the sets\/
$X_0$ and\/ $X_1$ also belong to $\pe\zeta$.
\ele
\noindent{\bft Proof\hspace{2mm} } The splitting of $X_e$ via an admissible $\Phi$ is equal to
the splitting of $X$ itself via the function $\Psi$ defined so
that $\Psi(0)=e$ and $\Psi(m+1)=\Phi(m)$.\qed
\begin{corollary}\TF\
\label{itspl}
If\/ $X\in\pe\zeta$ then\/
$X_\Phi[u]\in\pe\zeta$ for all\/ $u\in 2^{<\om}$.\qed
\end{corollary}
We now come to the principal point which was perhaps the only
reason for the introduction of shrinkability to the definition
of forcing conditions.
\bpro
\label{clop}
Assume that\/ $\zeta\in\IS,$ $X\in\pe\zeta,$ $X'\subseteq X$ is open in\/
$X,$ and\/ $x_0\in X'.$ There exists a clopen in\/ $X$ set\/
$X''\in\pe\zeta,$ $X''\subseteq X',$ containing $x_0$.
\epro
\noindent{\bft Proof\hspace{2mm} } Let $\Phi$ be a \dd\zeta admissible function. Then
$\ans{x_0}=X_\Phi[a]$ for some (unique) $a\in 2^\om.$ By
compactness there exists $m\in\om$ such that
$X''=X_\Phi[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m]\subseteq X'$.\qed
\newpage
\subsection{Finite iterations of splitting}
\label{ffin}
We shall exploit later the construction of sets in $\perf$ as
$X=\bigcap_{m\in\om}\bigcup_{u\in 2^m} X_u,$ where every $X_u$
belongs to $\perf.$ Each level $\ang{X_u:u\in 2^m}$ of the given
family of sets $X_u\in\perf$ should satisfy certain requirements
which resemble the state as if we had defined the sets by
iteration of splitting. This section introduces the requirements
and presents some related lemmas.
To specify the requirements which imply a good behaviour of the
sets $X_u$ with respect to projections, we need to determine, for
any pair of infinite binary sequences $u,\,v\in 2^{\<\om}$ of
equal length, the largest initial segment $\xi=\xi[u,v]$ such that
$X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ should coincide with $X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$.
\bdf
Let $\Phi$ be an \dd\bI admissible function. We define, for any
pair of finite sequences $u,\,v\in 2^m,$ $m\in\om,$ an initial
segment $\xi[u,v]=\xi_\Phi[u,v]\in \IS$ by induction as follows.\vspace{-1mm}
\begin{enumerate}
\item\label{xi0}
$\xi[\La,\La]=\bI$. ($\La$ is the empty sequence.)\vspace{-1mm}
\item\label{xieq}
$\xi[u{\mathbin{\kern 1.3pt ^\wedge}} d,v{\mathbin{\kern 1.3pt ^\wedge}} e]=\xi[u,v]$ provided $d=e\in\ans{0,1}$.\vspace{-1mm}
\item\label{xiOm}
Assume that $u$ and $v$ belong to $2^n,$
$d,\,e\in\ans{0,1}$ are different, and $\Phi(n)=\bald{i}\in\bI.$
Then $\xi[u{\mathbin{\kern 1.3pt ^\wedge}} d,v{\mathbin{\kern 1.3pt ^\wedge}} e]=\xi[u,v]\cap[\not\>\hspace{-2pt}\bald{i}]$.\vspace{-1mm}\qed
\end{enumerate}
\edf
\bdf
Let $\Phi$ be an \dd\bI admissible function.
A \dd\Phi{\it splitting system\/} of order $m$ is an indexed
family $\ang{X_u:u\in 2^{\<m}}$ of sets $X_u\in \perf$ such that\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){\rm{S-}\arabic{enumi}}
\item\label{spli}
$X_{u{\mathbin{\kern 1.3pt ^\wedge}} e}\subseteq \rbox{Spl}(X_u,\Phi(n),e)$ whenever
$u\in 2^n,$ $n<m,$ and $e\in\ans{0,1}$.\vspace{-1mm}
\item\label{prct}
$X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[u,v]=X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[u,v]$ for any pair of
$u,\,v$ of equal length $\<m$.\vspace{-1mm}
\item\label{aprct}
If $\bald{i}\in\bI,$ $\bald{i}\not\in\xi[u,v],$ then
$X_u\rsd{\<\bald{i}} \cap X_v\rsd{\<\bald{i}}=\emptyset$.\qed
\end{enumerate}
\edf
\noindent In particular it easily follows from lemmas \ref{spl'} and
\ref{spl} that for all $m$ and $X\in\perf$ the system defined by
$X_u=X_\Phi[u]$ for each $u\in 2^{\<m},$ is a \dd\Phi splitting
system.
We consider two ways how an existing splitting system can be
transformed to another splitting system. One of them treats
the case when we have to change one of sets to a smaller set,
the other one is an expansion to the next level.
It is assumed that a \dd\bI admissible function $\Phi$ is fixed
and $\xi[u,v]=\xi_\Phi[u,v]$.
\ble
\label{suz}
Assume that\/ $\ang{X_u:u\in 2^{\<m}}$ is a splitting system,
$u_0\in 2^m,$ and\/ $X\in\perf,$ $X\subseteq X_{u_0}.$ We re--define\/
$X_u$ by\/
$X'_u=X_u\cap (X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[u,u_0]\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI)$ for all\/ $u\in 2^m.$
Then the re--defined~\footnote
{\rm\ Notice that $X'_{u_0}=X$.}
family is again a splitting system.
\ele
\noindent{\bft Proof\hspace{2mm} } All sets $X'_u$ belong to $\perf$ by Lemma~\ref{apro}.
Therefore we have to check only condition~\ref{prct}. Thus let
$u,\,v\in 2^m,$ $\xi=\xi[u,v].$ We prove that
$X'_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X'_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$. Let in addition $\xi_u=\xi[u,u_0]$
and $\xi_v=\xi[v,u_0].$ Then
$$
X'_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\,\cap\,(X_0{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cap\xi_u)\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\xi)
\hspace{4mm}\hbox{and}\hspace{4mm}
X'_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\,\cap\,(X_0{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cap\xi_v)\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\xi)
$$
by Lemma~\ref{pro'}. Thus it remains to prove that
$\xi\cap\xi_u=\xi\cap\xi_v.$ Assume that on the contrary
$\bald{i}\in\xi\cap\xi_u$ but $\bald{i}\not\in\xi_v.$ The 1st assumption
implies $X_{u_0}\rsd{\<\bald{i}}=X_u\rsd{\<\bald{i}}=X_v\rsd{\<\bald{i}}$ by
condition~\ref{prct}, but the 2nd one implies that
$X_{u_0}\rsd{\<\bald{i}}\cap X_v\rsd{\<\bald{i}}=\emptyset,$
contradiction.\qed
\ble
\label{pand}
Every splitting system\/ $\ang{X_u:u\in 2^{\<m}}$ can be expanded
to the next level by adjoining appropriate sets\/ $X_{u'},$
$u'\in 2^{m+1}$.
\ele
\noindent{\bft Proof\hspace{2mm} } Let $\Phi(m)=\bald{i}\in\bI.$ We define
$X_{u{\mathbin{\kern 1.3pt ^\wedge}} e}=\rbox{Spl}(X_u,\bald{i},e)$ for all $u\in 2^m$ and $e=0,\,1.$ It
suffices to prove conditions \ref{prct} and \ref{aprct}.
Let $u'=u{\mathbin{\kern 1.3pt ^\wedge}} d$ and $v'=v{\mathbin{\kern 1.3pt ^\wedge}} e$ belong to $2^{m+1}$. We put
$Y=X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ where $\xi=\xi[u,v]$.\vspace{1mm}
{\it Case 1\/}: $\bald{i}\not\in\xi.$ Then by definition
$\xi=\xi[u',v']$ as well. Lemma~\ref{spl'}
immediately gives $X_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=Y=X_{v'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi.$ This proves
\ref{prct}. On the other hand, if $\bald{i}\not \in \xi$ then already
$X_u\rsd{\<\bald{i}}\cap X_v\rsd{\<\bald{i}}=\emptyset,$ and we have
\ref{aprct}.\vspace{1mm}
{\it Case 2\/}: $\bald{i}\in\xi$ and $d=e,$ say $d=e=0.$ Then again
$\xi=\xi[u{\mathbin{\kern 1.3pt ^\wedge}} d,v{\mathbin{\kern 1.3pt ^\wedge}} e]=\xi[u,v].$ We obtain
$X_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=\rbox{Spl}(Y,\bald{i},0)=X_{v'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ as required.\vspace{1mm}
{\it Case 3\/}. $\bald{i}\in\xi$ and $d\not=e,$ say $d=0,$ $e=1.$
Then $\xi'=\xi[u{\mathbin{\kern 1.3pt ^\wedge}} d,v{\mathbin{\kern 1.3pt ^\wedge}} e]=\xi\cap[\not\>\hspace{-2pt}\bald{i}],$ and we
obtain
$X_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'=\rbox{Spl}(Y,\bald{i},0){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'=\rbox{Spl}(Y,\bald{i},1){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'=
X_{v'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi',$ as required. (The middle equality follows from
Lemma~\ref{spl'} for $\zeta=\xi.$) To check~\ref{aprct} for some
$\bald{j},$ note
that if $\bald{j}\not\in\xi'$ then either already $\bald{j}\not\in\xi$ or
$\bald{j}\>\bald{i}.$ In the latter case we use Lemma~\ref{spl'} again
to obtain $\rbox{Spl}(Y,\bald{i},0)\rsd{\<\bald{i}}\cap \rbox{Spl}(Y,\bald{i},1)\rsd{\<\bald{i}}=
\emptyset.$ But $X_{u'}\rsd{\<\bald{i}}=\rbox{Spl}(Y,\bald{i},0)\rsd{\<\bald{i}}$ and
$X_{v'}\rsd{\<\bald{i}}=\rbox{Spl}(Y,\bald{i},1)\rsd{\<\bald{i}}$ because $\bald{i}\in\xi$ and
$Y=X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$.\qed
\newpage
\subsection{The forcing is homogeneous}
\label{hom}
The following theotem shows that the forcing $\pe\zeta$ is quite
homogeneous. This fact will be used in the proof of Theorem~\ref{ge}
in Section~\ref{seqge} but not earlier.
\bte
\label{thom}
Suppose that\/ $\zeta\in\IS.$ Let\/ $X,\,Y\in\pe\zeta.$ There exists
a homeomorphism\/ $F:X\,\lra\,Y$ such that\vspace{-1mm}
\begin{enumerate}
\item\label{h1} For all\/ $\xi\in\IS,$ $\xi\subseteq\zeta,$ and $x,\,y\in X,$
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi = y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ iff\/ $F(x){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi = F(y){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$.\vspace{-1mm}
\item\label{h2} If\/ $\xi\in\IS,$ $\xi\subseteq\zeta,$ and $X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$
then\/ $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=F(x){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ for all\/ $x\in X$.\vspace{-1mm}
\item\label{h3} For each\/ $X'\subseteq X$
we have\/ $X'\in\pe\zeta$ iff\/ $Y'=F{\hbox{\hspace{1pt}\rmt ''}} X'\in\pe\zeta$.\\
{\rm ($F{\hbox{\hspace{1pt}\rmt ''}} X=\ans{F(x):x\in X}$ is the image of $X$ via $F$.)}
\end{enumerate}
\ete
\noindent{\bft Proof\hspace{2mm} } Let us fix a \dd\zeta admissible function $\Phi.$ We recall
that the initial segment $\xi[u,v]=\xi_\Phi[u,v]\in\IS$ is defined
for $u,\,v\in 2^n$ by induction on $n$ in Section~\ref{ffin}. We
set $\xi[a,b]=\bigcap_n\xi[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n,b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n]$ for $a,\,b\in 2^\om;$
then $\xi[a,b]\in\IS$ as well.
We defined sets $X[u]=X_\Phi[u]\;\;(u\in 2^{<\om})$ in the end of
Section~\ref{splt}; all of them belong to $\pe\zeta$ by
Corollary~\ref{itspl}. Lemma~\ref{pand} implies:
\begin{enumerate}
\def\arabic{enumi}){{\hskip1pt\rm{s-}\arabic{enumi}\hskip1pt}}
\item\label{s1}\hspace{-1\mathsurround}
$X[u{\mathbin{\kern 1.3pt ^\wedge}} e]\subseteq\rbox{Spl}(X[u],\Phi(n),e)$ whenever
$u\in 2^n,$ $n<m,$ and $e\in\ans{0,1}$.\vspace{-1mm}
\item\label{s2}\hspace{-1\mathsurround}
$X[u]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[u,v]=X[v]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[u,v]$ for any pair of
$u,\,v$ of equal length $m\in\om$.\vspace{-1mm}
\item\label{s3}
If $\bald{i}\in\bI,$ $\bald{i}\not\in\xi[u,v],$ then
$X[u]\rsd{\<\bald{i}}\cap X[v]\rsd{\<\bald{i}}=\emptyset$.\qed
\end{enumerate}
Since $X,\,Y\in\pe\zeta,$ the sets $X[a]=\bigcap_nX[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n]$ and
$Y[a]=\bigcap_nY[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n]$ contain one point each, resp. $x_a$
and $y_a,$ for every $a\in 2^\om.$ It follows from \ref{s2}
and \ref{s3} that\vspace{-1mm}
\begin{itemize}
\item[$(\ast)$] $\hspace{-1\mathsurround} x_a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi[a,b]=x_b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi[a,b]$ \
for all \ $a,\,b\in 2^\om.$ \ \ If \ $\bald{i}\not\in\xi[a,b]$ \ then \
$x_a\rsd{\<\bald{i}}\not=x_b\rsd{\<\bald{i}}$.\vspace{-1mm}
\end{itemize}
It follows that $x_a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi= x_b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ iff $\xi\subseteq\xi[a,b]$ for
each pair of $a,\,b\in 2^\om$ and every $\xi\in\IS,\;\,\xi\subseteq\zeta;$
the same is true for $y_a,\,y_b.$
Taking $\xi=\zeta,$ we define a homeomorphism $F:X\,$onto$\,Y$ by
$F(x_a)=y_a$ for all $a\in 2^\om.$ For each
$\xi\in\IS,\;\,\xi\subseteq\zeta$ we obtain an associated homeomorphism
$F_\xi:X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\,$onto$\,Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ which satisfies\vspace{-1mm}
\begin{itemize}
\item[$(\dag)$]
$F_\xi(x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)=F(x){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ \ for all \ $x\in X$.\vspace{-1mm}
\end{itemize}
Let us prove that $F$ is as required. Item~\ref{h1}
of the theorem immediately follows from $(\dag).$ To verify
item~\ref{h2}, one easily proves that $X[u]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=Y[u]{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ for
all $u\in 2^{<\om}$ by induction on the length of $u,$ provided
$X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi.$ It remains to check item~\ref{h3}.
Let $X'\subseteq X,\;\,X'\in\pe\zeta;$ we have to prove that
$Y'=F{\hbox{\hspace{1pt}\rmt ''}} X'\in\pe\zeta$ as well.\vspace{1mm}
{\it We check requirement\/ \ref{perf1}\/}.
Thus let $\bald{i}\in\zeta$ and $y'\in Y'\rsd{<\bald{i}}.$ To prove that
$\hat Y=D_{Y'y'}(\bald{i})$ is a perfect set, let
$y'=F_{<\bald{i}}(x');\;x'\in X'\rsd{<\bald{i}}.$ Then the set
$\hat X=D_{X'x'}(\bald{i})$ is perfect. Let $\hat x\in\hat X;$ then
$\hat x=x(\bald{i})$ for some $x\in X'$ such that $x\rsd{<\bald{i}}=x'.$
One can define $y=F(x)$ and $\hat y=y(\bald{i})$ -- then
$y\in Y',$ $y\rsd{<\bald{i}}=y',$ and $\hat y\in \hat Y$ by $(\dag).$
It also follows from $(\dag)$ that in fact $\hat y$ depends
only on $\hat x$ but not on the entire $x;$ let
$\hat y=\hat F(\hat x).$ We conclude that $\hat F$ is a
homeomerphism $\hat X$ onto $\hat Y,$ as required.\vspace{1mm}
{\it We check requirement\/ \ref{oz}\/}. Suppose that
$\bald{i}\in\zeta$ and $Y^\ast\subseteq Y'$ is clopen in $Y';$ we have to
prove that the set $Y^\ast\rsd{<\bald{i}}$ is open in $Y'\rsd{<\bald{i}}.$
(This is more than \ref{oz} asserts, but here this is more
convenient.)
We put $X^\ast=F^{-1}(Y^\ast);$ then $X^\ast$ is clopen in
$X'$ because $F$ is a homeomorphism, and
$Y^\ast\rsd{<\bald{i}}=F_{<\bald{i}}{\hbox{\hspace{1pt}\rmt ''}} (X^\ast\rsd{<\bald{i}})$ by $(\dag);$
therefore it suffices to prove that
$X^\ast\rsd{<\bald{i}}$ is clopen in $X'\rsd{<\bald{i}}$.
Since each $X'[a]=X'_\Phi[a]\;\;(a\in 2^\om)$ is a singleton by
the shrinkability, there exists a number $n\in\om$ such that we
have $X'[u]\subseteq X^\ast$ or $X'[u]\cap X^\ast=\emptyset$ for every
$u\in 2^n.$ It follows from \ref{s2} and \ref{s3} above that
the projections $X'[u]\rsd{<\bald{i}}$ and $X'[v]\rsd{<\bald{i}}$ are either
equal or disjoint for each pair of $u,\,v\in 2^n.$ This easily
implies the required fact.\vspace{1mm}
{\it We check requirement\/ \ref{indep}\/}. Let
$\xi,\,\eta\in\IS,$ $\xi\cup\eta\subseteq\zeta,$ $y'\in Y'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$
$y''\in Y'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \eta,$ and
$y'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cap \eta)=y''{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cap \eta);$ we have to prove
that $y'\cup y''\in Y'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cup\eta).$ Using $(\dag),$ we
see that $y'=F_\xi(x')$ and $y''=F_\eta(x''),$ where
$x'\in X'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ $x''\in X'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \eta,$ and
$x'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cap \eta)=x''{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cap \eta).$ Then
$x'\cup x''\in X'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cup\eta)$ because $X'\in\pe\zeta.$ We
conclude that
$y'\cup y''=
F_{\xi\cup\eta}(x'\cup x'')\in Y'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} (\xi\cup\eta),$
as required.\vspace{1mm}
{\it We prove that $Y'$ is shrinkable\/}. Since $X'$ is shrinkable,
it suffices to verify that $Y'[u]=F{\hbox{\hspace{1pt}\rmt ''}} X'[u]$ for all
$u\in2^{<\om}.$ It follows from $(\dag)$ that the problem can be
reduced to one--dimentional setting.
Let $P$ be a perfect subset of ${\skri D}=2^\om.$ We set $P[\La]=P$ and
$P[u{\mathbin{\kern 1.3pt ^\wedge}} e]=\rbox{Spl}(P[u],e)$ for all $u\in 2^{<\om}$ and $e\in\ans{0,1}$
(see the end of Section~\ref{splt}); thus a perfect set $P[u]$ is
defined for
all $u\in 2^{<\om}.$ Obviously $P[a]=\bigcap_n P[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n]$ is a
singleton, say $p_a,$ for every $a\in 2^\om.$ Thus
$a\,\longmapsto\,p_a$ is a homeomorphism $2^\om$ onto $P$.
Suppose that $P,\,Q$ is a pair of perfect subsets of ${\skri D},$
with the associated homeomorphisms $a\longmapsto p_a$ and
$a\longmapsto q_a.$ One
defines a special homeomorphism $f=f_{PQ}:P\,$onto$\,Q$ by
$f(p_a)=q_a$ for all $a\in 2^\om.$ Now, the abovementioned
one--dimentional assertion
in the verification of shrinkability, is as follows:\vspace{-1mm}
\begin{itemize}
\item[$(\ddag)$] {\it prove that\/ $f{\hbox{\hspace{1pt}\rmt ''}} (P'[u])=(f{\hbox{\hspace{1pt}\rmt ''}} P')[u]$ for
every perfect\/ $P'\subseteq P$ and all\/ $u\in2^{<\om}$.}\vspace{-1mm}
\end{itemize}
We prove $f{\hbox{\hspace{1pt}\rmt ''}} (P'[u])=(f{\hbox{\hspace{1pt}\rmt ''}} P')[u]$ for a fixed perfect $P'\subseteq P$
by induction on the length of $u\in 2^{<\om}.$ The case $u=\La$ is
clear. We now suppose that $f{\hbox{\hspace{1pt}\rmt ''}} (P'[u])=Q'[u],$ where
$Q'=f{\hbox{\hspace{1pt}\rmt ''}} P',$ and prove
$f{\hbox{\hspace{1pt}\rmt ''}} (P'[u{\mathbin{\kern 1.3pt ^\wedge}} e])=Q'[u{\mathbin{\kern 1.3pt ^\wedge}} e]\,,$ $e=0,\,1$.
Let $s\in 2^{<\om}$ be the maximal sequence such that
$P'[u]\subseteq P[s].$ By definition $f=f_{PQ}$ maps $P[s]$ onto
$Q[s];$ by the assumption $f$ also maps $P'[u]$ onto
$Q'[u].$ We observe that $Q'[u]\subseteq Q[s]\,;$ moreover,
$Q'[u]\not\subseteq Q[s{\mathbin{\kern 1.3pt ^\wedge}} e]$ for any $e=0,\,1$ (otherwise we would
get $P'[u]\subseteq P[s{\mathbin{\kern 1.3pt ^\wedge}} e]\,,$ contradiction with the choice of $s$).
It follows that
$$
P'[u{\mathbin{\kern 1.3pt ^\wedge}} e]=P'[u]\cap P[s{\mathbin{\kern 1.3pt ^\wedge}} e]
\hspace{8mm}\hbox{and}\hspace{8mm}
Q'[u{\mathbin{\kern 1.3pt ^\wedge}} e]=Q'[u]\cap Q[s{\mathbin{\kern 1.3pt ^\wedge}} e]
$$
for $e=0,\,1.$
This implies $f{\hbox{\hspace{1pt}\rmt ''}} (P'[u{\mathbin{\kern 1.3pt ^\wedge}} e])=Q'[u{\mathbin{\kern 1.3pt ^\wedge}} e]\,,$ as required.\qed
\newpage
\subsection{Fusion technique}
\label{finf}
We prove here a version of {\it fusion lemma\/} which fits to this
form of iterated Sacks forcing.
An \dd\bI admissible function $\Phi$ is fixed; all notions of the
preceding section are treated in the sense of this function $\Phi$.
\bdf
An indexed family of sets $X_u\in\perf,$ $u\in 2^{<\om},$
is a {\it fusion sequence\/} if first, for every $m\in \om$
the subfamily $\ang{X_u:u\in 2^{\<m}}$ is a splitting system,
and second,\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){\rm{S-}\arabic{enumi}}
\setcounter{enumi}{3}
\item\label{di}
For any $\varepsilon>0$ there exists $m\in\om$ such that $\rbox{diam}\hspace{1pt} X_u<\varepsilon$
for all $u\in 2^m.$
(A Polish metric on $\can{\bI}$ is fixed.)\qed
\end{enumerate}
\edf
\bte
\label{fut}
Let\/ $\ang{X_u:u\in 2^{<\om}}$ be a fusion sequence. Then\/
$X=\bigcap_{m\in\om}\bigcup_{u\in 2^m} X_u$ belongs to
$\perf$.
\ete
\noindent{\bft Proof\hspace{2mm} } The intersection $\bigcap_m X_{a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m}$ contains a single
point $x_a\in X$ for any $a\in 2^\om={\skri D}$ by \ref{di}, and the
map $a\longmapsto x_a$ is continuous. Let us define
$\xi[a,b]=\bigcap_{m\in\om}\xi[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m,b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m].$ In particular
$\xi[a,b]=\bI$ iff $a=b.$ It follows from conditions \ref{prct} and
\ref{aprct} that\vspace{-1mm}
\begin{itemize}
\item[$(\ast)$] $\hspace{-1\mathsurround} x_a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi[a,b]=x_b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} \xi[a,b]$
for all $a,\,b\in 2^\om.$ If $\bald{i}\not\in\xi[a,b]$ then
$x_a\rsd{\<\bald{i}}\not=x_b\rsd{\<\bald{i}}.$\vspace{-1mm}
\end{itemize}
\noindent We check condition~\ref{perf1}. Let $\bald{i}\in \bI$ and
$y\in X\rsd{<\bald{i}},$ $p\in D=D_{Xy}(\bald{i}).$ For a fixed $k\in\om,$ we
shall find a point $q\in D,$ $q\not=p,$ which satisfies
$q{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} k=p{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} k.$ Let $x\in X$ be such that $p=x(\bald{i})$ and
$x\rsd{<\bald{i}}=y.$ Let $x=x_a;\;a\in 2^\om.$ First of all we fix
$m\in \om$ such that $x'(\bald{i}){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} k=x''(\bald{i}){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} k$ for all
$u\in 2^m$ and $x',\,x''\in X_u.$ Let $b\in 2^\om$ be defined by
$b(n)=a(n)$ for all $n\in\om$ with the exception of the first
$n>m$ such that $\Phi(n)=\bald{i},$ where we set $b(n)=1-a(n).$ Let
$q=x_b(\bald{i})$. Then\vspace{-1mm}
\begin{itemize}
\item[$1)$]
taking $u=a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m=b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m,$ we get $p{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} k= q{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} k$;\vspace{-1mm}
\item[$2)$] evidently $[<\hspace{-2pt}\bald{i}]\subseteq\xi[a,b],$ therefore
$x_b\rsd{<\bald{i}}=x_a\rsd{<\bald{i}}=y,$ so that $q\in D$;\vspace{-1mm}
\item[$3)$]\hspace{-1\mathsurround}
$q=x_b(\bald{i})\not=x_a(\bald{i})=p$ because $a(n)\not=b(n)$ for some
$n$ such that $\Phi(n)=\bald{i}$.\vspace{-1mm}
\end{itemize}
(In the last two items, we use $(\ast)$.)\vspace{1mm}
We check condition~\ref{oz} for the set $X.$ Let $\bald{i}\in \bI$ and
$G\subseteq{\skri D}$ be an open set. We define $X'=\ans{x\in X:x(\bald{i})\in G}$
and prove that $X'\rsd{<\bald{i}}$ is open in $X\rsd{<\bald{i}}$.
We can assume that $G$ is in fact clopen. Then by \ref{di}
there exists $m\in\om$ such that for every $u\in 2^m$ either
$X_u(\bald{i})\subseteq G$ or $X_u(\bald{i})\cap G=\emptyset.$ Let
$U=\ans{u\in 2^m:X_u(\bald{i})\subseteq G}.$ Notice that in accordance with
conditions \ref{prct} and \ref{aprct}, for any pair
$u,\,v\in2^m,$ either $X_u\rsd{<\bald{i}}=X_v\rsd{<\bald{i}}$ or
$X_u\rsd{<\bald{i}}\cap X_v\rsd{<\bald{i}}=\emptyset.$ Let $V$ be the set of
all $v\in 2^m$ such that $X_u\rsd{<\bald{i}}\cap X_v\rsd{<\bald{i}}=\emptyset$
for all $u\in U.$ There exists a clopen set $C\subseteq\can{<\bald{i}}$ which
separates $A=\bigcup_{u\in U}X_u\rsd{<\bald{i}}$ from
$B=\bigcup_{v\in V}X_v\rsd{<\bald{i}}.$ It remains to verify that
$X'\rsd{<\bald{i}}=X\rsd{<\bald{i}}\cap C$.
Let $x\in X',$ that is, $x=x_a\in X$ and $x(\bald{i})\in G.$ Then $u=
a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m\in U,$ therefore $x\rsd{<\bald{i}}\in A$ and $\in C.$ Let for
the converse $x=x_a\in X$ and $x\rsd{<\bald{i}}\in C;$ we have to find
$x'\in X'$ such that $x'\rsd{<\bald{i}}=x\rsd{<\bald{i}}.$ Notice that $v=
a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m$ does not belong to $V$ (although may not belong to $U$
either). Therefore $X_v\rsd{<\bald{i}}$ is equal to $X_u\rsd{<\bald{i}}$ for
some $u\in U.$ Let $b\in 2^\om$ be defined by $b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m=u$ and
$b(n)=a(n)$ for $n\>m.$ Then, since $X_v\rsd{<\bald{i}}=X_u\rsd{<\bald{i}},$
we obtain $X_{a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n}\rsd{<\bald{i}}=X_{b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n}\rsd{<\bald{i}}$ for all
$n>m.$ Therefore, $x'=x_b$ satisfies $x'\rsd{<\bald{i}}=x\rsd{<\bald{i}}.$ On
the other hand, $x'\in X'$ because $b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m=u\in U$.\vspace{1mm}
We check \ref{indep}. Let $\xi,\,\eta\in\IS,$ and points
$x'=x_{a'},\,x''=x_{a''}$ of $X$ be such that
$x'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cap\eta)=x''{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi\cap\eta);$ then in particular
$\xi\cap\eta\subseteq\xi[a',a'']$ by $(\ast)$. We have to find
$x=x_a\in X$ satisfying $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=x'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta=x''{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$.
To obtain the required $a\in 2^\om,$ we define the values
$a(m)\in\ans{0,1}$ using induction on $m.$ Assume that we have
defined $a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m,$ and define $a(m)$. Let $\Phi(m)=\bald{i}\in\bI$.\vspace{1mm}
{\it Case 1\/}: $\bald{i}\not\in\xi\cup\eta.$ We define $a(m)$
arbitrarily, say $a(m)=0$ in this case.
{\it Case 2\/}: $\bald{i}\in\xi\setminus\eta$ or
$\bald{i}\in\eta\setminus\xi.$ We put resp. $a(m)=a'(m)$ or
$a(m)=a''(m)$.
{\it Case 3\/}: $\bald{i}\in\xi\cap\eta.$ Then $a'(m)=a''(m)$ since
otherwise we would have $\bald{i}\not\in\xi[a',a''].$ We put
$a(m)=a'(m)=a''(m)$.\vspace{1mm}
This definition implies $\xi\subseteq\xi[a,a']$ and
$\eta\subseteq\xi[a,a''].$ Then $x_a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=x_{a'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and
$x_a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta=x_{a''}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ by $(\ast),$ as required.
Finally to check shrinkability we note that $X_\Phi[u]=
X\cap X_u$ (can be easily proved by induction), so that the
property follows from~\ref{di}.\qed
\begin{corollary}\TF\
\label{prebor}
Let\/ $\xi\in\IS,$ $X\in\pe\xi,$ and\/ $C_m\subseteq\can\xi$ be closed
for each $m\in\om.$ There exists\/ $Y\in\pe\xi,$ $Y\subseteq X$ such
that\/ $C_m\cap Y$ is clopen in\/ $Y$ for every\/ $m$.
\end{corollary}
\noindent{\bft Proof\hspace{2mm} } We can assume that $\xi=\bI$ (if not replace $C_m$ by
$C_m\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI$). It follows from Proposition~\ref{clop} that for
any $m$ and any $X'\in\perf$ there exists $Y'\in\perf,$ $Y'\subseteq X',$
such that either $Y'\subseteq C_m$ or $Y'\cap C_m=\emptyset.$ Therefore
we can define, using lemmas \ref{suz} and \ref{pand}, a fusion
sequence $\ang{X_u:u\in 2^{<\om}}$ of sets $X_u\in\perf$ such that
$X_\La=X$ and either $X_u\subseteq C_m$ or $X_u\cap C_m=\emptyset$
whenever
$u\in 2^m.$ Let $Y=\bigcap_{m\in\om}\bigcup_{u\in 2^m}X_u$.\qed
\begin{corollary}\TF\
\label{bor}
Assume that\/ $\xi\in\IS,$ $X\in\pe\xi,$ and\/ $B\subseteq\can\xi$
is a set of a finite Borel level. There exists\/ $Y\in\pe\xi,$
$Y\subseteq X$ such that either\/ $Y\subseteq B$ or\/ $Y\cap B=\emptyset$.
\end{corollary}
\noindent{\bft Proof\hspace{2mm} }\footnote
{\rm\ In fact this is true for all Borel sets $B$ but needs
more elaborate reasoning.} \
Let $B$ be defined by a finite level Borel scheme (countable
unions plus countable intersections) from closed sets $
C_m,\;m\in\om.$ The preceding corollary shows that there exists
$X'\in\pe\xi,$ $X'\subseteq X$ such that every $X'\cap C_m$ is clopen
in $X'.$ Thus the Borel level can be reduced.\qed\vspace{3mm}
The results already obtained are in fact sufficient to prove the
first part of Theorem~\ref{m}. However to handle the degrees of
constructibility in \dd\perf generic extensions we need to conduct a
more detailed analysis concentrated on continuous functions.
\newpage%
\subsection{Reducibility of continuous functions}
\label{cont}
This section provides analysis of the behaviour of continuous
functions defined on sets in $\perf$ from the point of view of
certain reducibility.
\bdf
For $\xi\subseteq\bI,$ $\cnt\xi$ is the set of all continuous
functions $F:\can\xi\;\lra\;\om^\om.$ We put $\cont=\cnt\bI.$
Let $F\in\cont$.\vspace{-1mm}
\begin{enumerate}
\item\label{redu}\hspace{-1\mathsurround}
$F$ is {\it reducible\/} to
$\xi\in\IS$ on a set $X\subseteq\can{\bI}$ if for all $x,\,y\in X$ such that
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ we have $F(x)=F(y)$.\vspace{-1mm}
\item\label{capt}\hspace{-1\mathsurround}
$F$ {\it captures\/} $\bald{i}\in\bI$ on $X$
if for all $x,\,y\in X$ such that $F(x)=F(y)$ we have
$x(\bald{i})=y(\bald{i})$.\qed
\end{enumerate}
\edf
\brem
\label{redr}
It follows from the compactness of the spaces we consider, that
if $X$ is closed then in item~\ref{redu} there exists a function
$H\in\cnt\xi$ such that $F(x)=H(x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)$ for all $x\in X,$ while
in item~\ref{capt} there exists a continuous function
$E:{\skri D}\;\lra\;{\skri D}$ such that $x(\bald{i})=E(F(x))$ for all $x\in X$.\qed
\erem
Theorem~\ref{m} contains four items, \ref{m1+} through \ref{m3},
concerning \dd{\bbb{\bf M}\bbb} constructibility of reals in the extension. We
shall obtain those assertions as corollaries of the following
theorem.
\bte
\label{mm}
Assume that\/ $\xi\in\IS,$ $X\in\perf,$ $F\in\cont.$ Then\vspace{-1mm}
\begin{enumerate}
\item\label{mm1+}
If\/ $\bald{i},\,\bald{j}\in\bI$ and\/ $\bald{i}<\bald{j}$ then
there exists\/ $X'\in\perf,\;\;X'\subseteq X,$ such that the
co-ordinate function $C_\bald{j}$ defined on\/ $\can{\bI}$ by
$C_\bald{j}(x)=x(\bald{j})$ captures\/ $\bald{i}$ on $X'$.\vspace{-1mm}
\item\label{mm2}
If\/ $\bald{i}\in\bI\setminus\xi$ and\/ $F$ is reducible to\/
$\xi$ on\/ $X$ then\/ $F$ does not capture\/ $\bald{i}$ on\/
$X$.\vspace{-1mm}
\item\label{mm2+}
Suppose that\/
$F$ satisfies the property that for all\/ $X'\in\perf,$
$X'\subseteq X,$ and\/ $\bald{i}\in\xi$ there exists\/ $X''\in\perf,$
$X''\subseteq X'$ such that\/ $F$ captures\/ $\bald{i}$ on\/ $X''.$ Then there
exists\/ $Y\in\perf,$ $Y\subseteq X$ such that\/ $F$ captures each\/
$\bald{i}\in\xi$ on\/ $Y$.\vspace{-1mm}
\item\label{mm3}
There exists\/ $X'\in\perf,$ $X'\subseteq X$ satisfying exactly one from
the following two assertions$:$\hspace{4mm}
\begin{minipage}[t]{0.6\textwidth}
{\rm(a)} $F$ is reducible to\/ $\xi$ on\/ $X'$, \hfill
or\vspace{2mm}
{\rm(b)} $F$ captures some\/ $\bald{i}\in\bI\setminus\xi$ on\/ $X'$.
\end{minipage}
\hfill $\,$
\end{enumerate}
\ete
\noindent{\bft Proof\hspace{2mm} } We begin from a few technical lemmas, then come
to the theorem.
\ble
\label{inter}
If\/ $F$ is reducible to both\/ $\xi$ and\/ $\eta$ on\/
$X\in\perf$ then\/ $F$ is reducible to\/ $\zeta=\xi\cap\eta$ on
$X$.
\ele
\noindent{\bft Proof\hspace{2mm} } Let, on the contrary, $x,\,y\in Y$ satisfy $x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\zeta=
y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\zeta$ but $F(x)\not=F(y).$ Then by property~\ref{indep}
of $X$ there exists $z\in X$ such that $z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and
$z{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta.$ We obtain $F(x)=F(z)=F(y),$
contradiction.\qed
\ble
\label{l1}
Assume that\/ $\xi\in\IS,$ $\bald{i}\in\bI\setminus\xi,$ the sets
$X_1$ and\/ $X_2$ belong to $\perf,$ and\/ $X_1{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X_2{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi.$
Then either $F$ is reducible to $\xi$ on $X_1\cup X_2$ or there
exist sets\/ $X_1',\,X_2'\in\perf,$ $X_1'\subseteq X_1$ and\/
$X_2'\subseteq X_2,$ such that again\/ $X'_1{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X'_2{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ and\/
$F{\hbox{\hspace{1pt}\rmt ''}} X'_1\cap F{\hbox{\hspace{1pt}\rmt ''}} X'_2=\emptyset$.~\footnote
{\rm\ We recall that $F{\hbox{\hspace{1pt}\rmt ''}} X$ is the image of $X$ via $F$.}
\ele
\noindent{\bft Proof\hspace{2mm} } We suppose that $F$ is not reducible to $\xi$ on
$X_1\cup X_2$ and prove the ``or'' alternative. By the
non--reducibility assumption there exist points
$x_1,x_2\in X_1\cup X_2$ such that $x_1{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi= x_2{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and
$F(x_1)\not=F(x_2).$ By the continuity of $F$ there exist
clopen neighbourhoods $U_1$ and $U_2$ of $x_1$ and $x_2$ such that
$F{\hbox{\hspace{1pt}\rmt ''}} U_1\cap F{\hbox{\hspace{1pt}\rmt ''}} U_2=\emptyset.$ Proposition~\ref{clop} gives
a set $X''_1\in\perf,$ $X''_1\subseteq X_1\cap U_1,$ containing $x_1$.
By Lemma~\ref{apro} the set
$X''_2=X_2\cap (X''_1{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI)$ belongs to $\perf,$ and
contains $x_2$ since $x_1{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=x_2{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi.$ By
Proposition~\ref{clop} again, there exists a set $Z_2\in\perf,$
$Z_2\subseteq X''_2\cap U_2.$ Now putting
$Z_1=X''_1\cap (Z_2{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI)$ we get sets
$Z_1,\,Z_2\in\perf$ such that $Z_1\subseteq X_1,$ $Z_2\subseteq X_2,$
$Z_1{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=Z_2{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and
$F{\hbox{\hspace{1pt}\rmt ''}} Z_1\cap F{\hbox{\hspace{1pt}\rmt ''}} Z_2=\emptyset$.\qed
\ble
\label{l2}
If\/ $F\in\cont,$ $X\in\perf,$ and\/ $\bald{i}\in\bI,$ then there
exists\/ $X'\in\perf,$ $X'\subseteq X$ such that either\/ $F$ is
reducible to\/ $[\not\>\hspace{-2pt}\bald{i}]$ on\/ $X',$ or\/ $F$
captures\/ $\bald{i}$ on $X'$.
\ele
\noindent{\bft Proof\hspace{2mm} } Let us assume that a set $X'$ of the ``either'' type does
not exist. To prove the existence of $X'$ of the ``or'' type,
we fix an \dd\bI admissible function $\Phi$ and put
$\xi[u,v]=\xi_\Phi[u,v]$ for every pair of finite sequences
$u,\,v\in 2^{<\om}$ of equal length. The notions of splitting
system and fusion sequence are understood in the sense of $\Phi$.
Using Lemma~\ref{l1} and Proposition~\ref{clop}, we define a
fusion sequence $\ang{X_u:u\in 2^{<\om}}$ with $X_\La=X,$
satisfying the following condition:\vspace{-1mm}
\begin{itemize}
\item [$(\star)$] If $m\in \om$ and $u,\,v\in 2^m$ then either
$(1)$ $F$ is reducible to $\xi[u,v]$ on the set $X_u\cup X_v$, \ or
$(2)$ $F{\hbox{\hspace{1pt}\rmt ''}} X_u\cap F{\hbox{\hspace{1pt}\rmt ''}} X_v=\emptyset$.
\vspace{-1mm}
\end{itemize}
Indeed we put $X_\La=X,$ as indicated. Assume that sets $X_u,$
$u\in 2^m,$ have been defined, and $\bald{i}=\Phi(m).$ We first set
$Y_{u{\mathbin{\kern 1.3pt ^\wedge}} e}=\rbox{Spl}(X_u,\bald{i},e)$ for all $u\in 2^m$ and $e=0,\,1,$
obtaining a splitting system $\ang{Y_{u'}:u'\in 2^{m+1}}$ (see
Lemma~\ref{pand}). At the next step we consider consequtively
all pairs $u',v'\in 2^{m+1}$ and reduce sets $Y_{u'}$ and
$Y_{v'}$ using Lemma~\ref{l1} for $\xi=\xi[u,v].$ Let $X'_{u'}$
and $X'_{v'}$ be the pair of sets we obtain; in particular
$X'_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X'_{v'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ and either $F$ is reducible
to $\xi$ on $X'_{u'}\cup X'_{v'}$ or
$F{\hbox{\hspace{1pt}\rmt ''}} X'_{u'}\cap F{\hbox{\hspace{1pt}\rmt ''}} X'_{v'}=\emptyset$.
Then we set
$Z_{w'}=X'_{w'}\cap (X'_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[w',u']\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI),$ so that
$\ang{Z_{w'}:w'\in 2^{m+1}}$ is a splitting system by
Lemma~\ref{suz}. It is essential that since
$X'_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X'_{v'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ we have $X'_{v'}\subseteq Z_{v'}.$
This allows to repeat the reduction: let
$Y'_{w'}=Z_{w'}\cap (X'_{v'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[w',u']\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI),$ which
gives again a splitting system of sets such that
$Y'_{u'}=X'_{u'}$ and $Y'_{v'}=X'_{v'}.$ This procedure eliminates
the particular pair of $u',\,v'\in 2^{m+1},$ as required.
Then $X'=\bigcap_m\bigcup_{u\in 2^m}X_u$ belongs to $\perf.$ We
prove that $X'$ is as required, that is, $F$ captures $\bald{i}$ on
$X'.$ Assume that, on the contrary, there exists a pair of points
$x,\,y\in X'$ such that $F(x)=F(y)$ but $x(\bald{i})\not=y(\bald{i}).$ Let
$x=x_a$ and $y=x_b$ (see the beginning of the proof of
Theorem~\ref{fut}), $a,\,b\in 2^\om.$ Then
$\bald{i}\not\in\xi[a,b]=\bigcap_m\xi[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m,b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m]$ (see assertion
$(\ast)$ in the proof of Theorem~\ref{fut}). Let $m$ be the least
among those satisfying $\bald{i}\not\in\xi=\xi[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m,b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m].$ Then
$\xi\subseteq[\not\>\hspace{-2pt}\bald{i}],$ so that the case $(1)$ in $(\star)$ is
impossible for $u=a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m$ and $v=b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m.$ Therefore
$F{\hbox{\hspace{1pt}\rmt ''}} X_u\cap F{\hbox{\hspace{1pt}\rmt ''}} X_v=\emptyset,$ contradiction with the choice
of $x$ and $y$ because $x\in X_u,$ $y\in X_v$.\qed\vspace{4mm}
We are already equipped enough to handle different items of
Theorem~\ref{mm}.\vspace{1mm}
{\it Item \ref{mm2}\/}. Thus suppose that $F$ is reducible to
$\xi$ on\/ $X$ and, on the contrary, $F$ does capture some
$\bald{i}\in\bI\setminus\xi$ on $X.$ Then the co-ordinate function
$C_\bald{i}(x)=x(\bald{i})$ is itself reducible to $\xi$ on $X.$ Since
$\bald{i}$ does not belong to $\xi,$ and on the other hand $C_\bald{i}$ is
obviously reducible to $[\<\hspace{-2pt}\bald{i}],$ we conclude that $C_\bald{i}$ is also
reducible to $[<\hspace{-2pt}\bald{i}]$ on $X$ by Lemma~\ref{inter} But this
clearly contradicts requirement~\ref{perf1}.\vspace{1mm}\vom
{\it Item \ref{mm1+}\/}. Otherwise, by Lemma~\ref{l2} $C_\bald{j}$ is
reducible to\/ $\xi=[\not\>\hspace{-2pt}\bald{i}]$ on some $X'\in\perf,$
$X'\subseteq X,$ contradiction with the already proved
item~\ref{mm2}.\vspace{1mm}\vom
{\it Item \ref{mm2+}\/}. Arguing as in the proof of Lemma~\ref{l2},
we get a fusion system $\ang{X_u:u\in 2^{<\om}}$ such that
$X_\La\subseteq X$ and $(\star)$ holds. We prove that the set
$Y=\bigcap_n\bigcup_{u\in 2^m} X_u$ is as required. Suppose that on
the contrary $x,\,y\in Y$ satisfy $F(x)=F(y);$ we have to prove that
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi.$ By definition, $\ans{x}=\bigcap_n X_{a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n}$
and $\ans{y}=\bigcap_n X_{b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} n}$
for certain (unique) $a,\,b\in 2^\om.$ It suffices to verify
that $\xi\subseteq\xi[a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m,b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m]$ for all $m$.
Assume that on the contrary $\xi\not\subseteq\xi[u,v],$ where $u=a{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m$
and $v=b{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}} m$ for some $m.$ We assert that the case $(1)$ of
$(\star)$ cannot happen. Indeed otherwise in particular $F$ is
reducible to $\xi[u,v]$ on a set $X'=X_u\subseteq X.$ Take an arbitrary
$\bald{i}\not\in\xi.$ By the assumption of item~\ref{mm2+} $F$ captures $\bald{i}$
on a set $X''\in\perf,$ $X''\subseteq X'.$ Thus the co-ordinate function
$C_\bald{i}$ is reduced to $\xi[u,v]$ on $X''$ -- contradiction with the
already proved item~\ref{mm2}. Thus we have case $(2)$ of $(\star),$
that is, $F{\hbox{\hspace{1pt}\rmt ''}} X_u\cap F{\hbox{\hspace{1pt}\rmt ''}} X_v=\emptyset.$
But this contradicts the assumption $F(x)=F(y)$.\vspace{1mm}\vom
{\it Item \ref{mm3}\/}. We fix $\Phi$ and put $\xi[u,v]=\xi_\Phi[u,v]$
as in the begining of the proof of Lemma~\ref{l2}. Assume that a set
$X'\in\perf$ of type (b) of item~\ref{mm3} does not exist.
Then by Lemma~\ref{l2} if
$\bald{i}\in\bI\setminus\xi$ then every set $Y\in\perf,$ $Y\subseteq X$
contains a subset $Z\in \perf$ such that $F$ is reducible to
$[\not\>\hspace{-2pt}\bald{i}]$ on $Z.$ Using lemmas \ref{suz} and \ref{pand} we
obtain a fusion sequence $\ang{X_u:u\in 2^{<\om}}$ such that
$X_\La\subseteq X$ and $F$ is reducible to $[\not\>\hspace{-2pt}\Phi(m)]$ on
$X_u$ whenever $u\in 2^m$ and $\Phi(m)\not\in\xi.$ Then
$X'=\bigcap_m\bigcup_{u\in 2^m}X_u\in\perf.$ We prove that $X'$ is
a set of type (a), that is, $F$ is reducible to $\xi$ on $X'$.
Let us define $\xi_m\in\IS$ by induction on $m$ so that
$\xi_0=\bI$ and
$$
\xi_{m+1}=\left\{
\begin{array}{ccl}
\xi_m & - & \hbox{in the case }\, \Phi(m)\in\zeta;\\[3mm]
\xi_m\cap [\not\>\hspace{-2pt}\Phi(m)] & - & \hbox{in the case }\,
\Phi(m)\not\in\zeta
\end{array}
\right.
$$
for all $m.$ Notice that then $\xi_m\subseteq\xi[u,v]$ whenever
$u,\,v\in 2^m$ satisfy $\xi\subseteq\xi[u,v]$.\vspace{3mm}
\noindent
{\bf Assertion} \ {\it For any\/ $m,$ $F$ is reducible to\/
$\xi_m$ on\/ $X_m=\bigcup_{u\in 2^m}X_u$.}\vspace{4mm}
\noindent
{\bf Proof} \ of the assertion.
We argue by induction on $m.$ Assume that, on the contrary, there
exist $u,\,v\in 2^{m+1}$ and $x\in X_u,$ $y\in X_v$ such that
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_{m+1}=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_{m+1},$ but $F(x)\not= F(y).$ Let
$u=u'{\mathbin{\kern 1.3pt ^\wedge}} d,$ $v=v'{\mathbin{\kern 1.3pt ^\wedge}} e,$ where $u',\,v'\in 2^m$ and
$d,\,e\in\ans{0,1}.$
We have $\xi_{m+1}\subseteq\xi[u,v]$ by \ref{prct} and
\ref{aprct}, therefore $\xi\subseteq\xi[u,v]$ because every set
$\xi_n$ includes $\xi.$ This implies $\xi\subseteq\xi[u',v'].$ It
follows (see above) that $\xi_m\subseteq\xi[u',v']$.
The nontrivial case is the case when $\bald{i}=\Phi(m)\not\in\xi$ since
otherwise $\xi_{m+1}=\xi_m$ and we can use the inductive
hypothesis. Then $\xi_{m+1}=\xi_m\cap [\not\>\bald{i}]$.
We assert that there exists $x'\in X_{u'}$ such that\vspace{-1mm}
\begin{itemize}
\item $x'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_m=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_m$\hspace{5mm} and\hspace{5mm}
$x'\rsd{\not\>\bald{i}}=x\rsd{\not\>\bald{i}}$.\vspace{-1mm}
\end{itemize}
Indeed, first $x\rsd{\not\>\bald{i}}\in X_{u'}\rsd{\not\>\bald{i}}.$ Second,
since $\xi_m\subseteq\xi[u',v'],$ we obtain
$X_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_m=X_{v'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_m$ by \ref{prct}. Therefore
$y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_m\in X_{u'}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi_m.$ Finally, using the fact that
$\xi_{m+1}=\xi_m\cap [\not\>\hspace{-2pt}\bald{i}]$ we conclude that
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi_m\cap[\not\>\hspace{-2pt}\bald{i}])=y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}(\xi_m\cap[\not\>\hspace{-2pt}\bald{i}])$
by the choice of $x$ and $y.$ Property~\ref{perf1} of $X_{u'}$
then implies the existence of a point $x'\in X_{u'}$ satisfying
$(\bullet)$.
To end the proof of the assertion, notice that $F(x')=F(y)$ by
the inductive hypothesis while $F(x')=F(x)$ by the choice of the
fusion sequence of sets $X_u$.\qed\vspace{4mm}
We continue the proof of item~\ref{mm3} of Theorem~\ref{mm}.
It follows from the
assertion that $F$ is reducible to every $\xi_m$ on $X'.$ This
allows to conclude that $F$ is also reducible to $\xi$ on $X'.$
Indeed assume that on the contrary $x,\,y\in X'$ and
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi= y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ but $F(x)\not=F(y).$ By the continuity of
$F$ there exist $m\in\om$ and $u,\,v\in 2^m$ such that $x\in X_u,$
$y\in X_v,$ and $F{\hbox{\hspace{1pt}\rmt ''}} X_u\cap F{\hbox{\hspace{1pt}\rmt ''}} X_v=\emptyset.$ Notice that
then ${X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\cap X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\not=\emptyset},$ therefore
$\xi\subseteq\xi[u,v]$ by {\ref{aprct}}.
This implies $\xi\subseteq\xi_m\subseteq\xi[u,v],$ as above. Therefore $F$ is
reducible to $\xi[u,v]$ on $X',$ contradiction with the equality
$F{\hbox{\hspace{1pt}\rmt ''}} X_u\cap F{\hbox{\hspace{1pt}\rmt ''}} X_v=\emptyset,$ because
$X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[u,v]=X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi[u,v]$ by condition \ref{prct}.\qed
\newpage
\subsection{Proof of the theorem: the countable case}
\label{re}
This section starts the proof of Theorem~\ref{m} in the case when the
``length'' $\bI$ of iteration is countable.
Thus let ${\bbb{\bf M}\bbb}$ be a countable
transitive model of $\ZFC,$ $\bI\in {\bbb{\bf M}\bbb}$ be a countable in ${\bbb{\bf M}\bbb}$
partially ordered set.\vspace{2mm}
\noindent{\bf The forcing}\\[2mm]
We consider $\perfm=(\perf)^{\bbb{\bf M}\bbb}$ as a forcing
notion ($X\subseteq Y$ means that $X$ is a stronger condition). Notice
that every set in $\mathord{{\rm I}\hspace{-2.5pt}{\rm P}}$ is then a countable subset in the
universe. However we can transform it to a perfect set by the
closure operation: the topological closure $X^\#$ of a set
$X\in \perfm$ will then satisfy the definition of $\perf$ from the
point of view of the universe.\vspace{2mm}
\noindent{\bf The model}\\[2mm]
Let $G\subseteq\perfm$ be a \dd\perfm generic
ultrafilter over ${\bbb{\bf M}\bbb}.$ Then the intersection
$\Pi=\bigcap_{X\in G}X^\#$ is a singleton (this easily follows
from Proposition~\ref{clop}). Let ${\bbsp\bbox{x}\bbsp}\in\can{\bI}$ be the unique
element of $\Pi;$ thus ${\bbsp\bbox{x}\bbsp}$ is a function from $\bI$ to reals. As
usual in this case the generic extension ${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[G]$ is equal
to ${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}]$.
We define ${\bbsp\bbox{a}\bbsp}_\bald{i}={\bbsp\bbox{x}\bbsp}(\bald{i})$ for all $\bald{i}\in\bI,$ so that
${\bbsp\bbox{x}\bbsp}=\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\bI}$.
\bpro
\label{counta}
The model\/ ${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}]$ satisfies the
requirements of Theorem~\ref{m}.
\epro
The proof of this proposition is the content of this section.
First we prove the cardinality
preservation property and an important technical theorem which
will allow to study reals in the extension using continuous
functions in the initial model.
\bte
\label{card}
$\aleph_1^{\bbb{\bf M}\bbb}$ and any cardinal greater than\/
$2^{\aleph_0}$ in\/ ${\bbb{\bf M}\bbb}$ remain cardinals in ${\bbb{\bf N}\bbb}$.
\ete
\noindent{\bft Proof\hspace{2mm} } Let ${\underline f}$ be a name of a function defined on $\om$ in the
language of forcing. Using in ${\bbb{\bf M}\bbb}$ lemmas \ref{suz} and
\ref{pand} we obtain a fusion sequence $\ang{X_u:u\in 2^{<\om}}$
of sets $X_u\in\perfm$ such that $X_\La\subseteq X_0$ (where $X_0\in\perfm$
is a given condition) and for all $m,$ every $X_u,\,\,u\in 2^m,$
decides the value of ${\underline f}(m).$ Then
$X=\bigcap_m\bigcup_{u\in 2^m}X_u\in\perfm,$ $X\subseteq X_0,$ and $X$
forces that the range of ${\underline f}$ is a subset of a countable in
${\bbb{\bf M}\bbb}$ set.\qed\vspace{4mm}
\noindent{\bf Continuous functions}\\[2mm]
We put $\cntm\xi=(\cnt\xi)^{\bbb{\bf M}\bbb}$ and $\contm=(\cont)^{\bbb{\bf M}\bbb}.$
It is a principal property of several forcing notions
(including Sacks forcing) that reals in the generic extension can
be obtained by application of continuous functions (having a code)
in the ground model, to the generic object. As we shall prove this
is also a property of the Sacks iteration considered here.
Obviously every $F\in\cntm\xi$ is a countable subset
of $\can\xi\times \om^\om$ in the universe, but since the domain of
$F$ in ${\bbb{\bf M}\bbb}$ is the compact set $\can\xi,$ $F^\#$ is a continuous
function mapping $\can\xi$ into the reals in the universe.
\bte
\label{repr}
Let\/ $\xi\in\IS$ and $c\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]\cap\om^\om.$ There
exists a function\/ $H\in\cntm\xi$ such that\/
$c=H^\#({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)$.~\footnote
{\rm\ Obviously this equality is absolute for any model containing
all of $c,\,{\bbsp\bbox{x}\bbsp},\,H$.}
\ete
\noindent{\bft Proof\hspace{2mm} } Let ${\underline c}$ be a name for $c$ containing an explicit
absolute construction of $c$ from ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and some parameter
$p\in{\bbb{\bf M}\bbb}.$
{\it We argue in ${\bbb{\bf M}\bbb}.$}
It follows from lemmas \ref{pro'} and
\ref{apro} that forcing of properties of ${\underline c}$ is reduced to
$\xi$ in the sense that if $X$ forces ${\underline c}(m)=k$ then
$X^-=X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI$ forces ${\underline c}(m)=k,$ too.
Therefore, given $X_0\in\perf,$ we can define a fusion sequence
$\ang{X_u:u\in 2^{<\om}}$ of sets $X_u\in\perf$ such that
$X_\La\subseteq X_0$ and, for all $m\in\om$ and $u\in 2^m,$ $X_u$
decides the value of ${\underline c}(m),$ say, forces that
${\underline c}(m)=\al(u),$ where $\al\in{\bbb{\bf M}\bbb}$ maps $2^{<\om}$ into $\om,$
so that if $X_u{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=X_v{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ then $\al(u)=\al(v).$ The
function $F'$ defined on
$X=\bigcap_m\bigcup_{u\in 2^m}X_u\in\perf$ by the property:
$F'(x)(m)=k$ iff $\al(u)=k,$ -- for all $m$ and all $u\in 2^m$
such that $x\in X_u,$ is continuous and reducible to $\xi$ on
$X.$ Therefore $F'$ can be expanded to a function $F\in\cont$
reducible to $\xi$ on $\can{\bI}.$ In this case (see
Remark~\ref{redr}) there exists a function $H\in\cnt\xi$ such
that $F(x)=H(x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)$ for all $x\in\can{\bI}.$ Then $X$ forces that
${\underline c}=F^\#({\bbsp\bbox{x}\bbsp})=H^\#({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)$.\qed\vspace{3mm}
This theorem practically reduces properties of reals in
\dd\perfm generic extensions to properties of continuous functions
in the ground model.
To demonstrate how Theorem~\ref{repr} works we prove statements
\ref{m1+} through \ref{m3} of Theorem~\ref{m} for the model
${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[G]$ of consideration.\vspace{3mm}
\noindent{\bf Proof of statement \ref{m1+} of Theorem~\ref{m}}\\[2mm]
Thus let $\bald{i},\,\bald{j}\in\bI,\;\;\bald{i}<\bald{j};$ we have to prove that
${\bbsp\bbox{a}\bbsp}_\bald{i}\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{a}\bbsp}_\bald{j}].$ It follows from Theorem~\ref{mm}
(item~\ref{mm1+}) that there exists a condition $X\in G$ such that
the following is true in ${\bbb{\bf M}\bbb}:$ the co-ordinate function $C_\bald{j}$
captures $\bald{i}$ on $X.$ In other words, in ${\bbb{\bf M}\bbb}$ there exists a
continuous function $H:{\skri D}\,\lra\,{\skri D}$ such that
$x(\bald{i})=H(x(\bald{j}))$ for all $x\in X.$ It follows that
$x(\bald{i})=H^\#(x(\bald{j}))$ for all $x\in X^\#$ is also true in ${\bbb{\bf N}\bbb}.$
($H^\#$ is the topological closure of $H$ as a subset of ${\skri D}^2$.)
Therefore ${\bbsp\bbox{a}\bbsp}_\bald{i}=H^\#({\bbsp\bbox{a}\bbsp}_\bald{j})\in {\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[G]$.\vspace{3mm}
\noindent{\bf Proof of statement \ref{m2} of Theorem~\ref{m}}\\[2mm]
Let $\xi\in {\bbb{\bf M}\bbb}$ be an
initial segment of $\bI,$ and $\bald{i}\in\bI\setminus\xi;$ we have
to prove that ${\bbsp\bbox{a}\bbsp}_\bald{i}\not\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi].$ Assume that on the
contrary ${\bbsp\bbox{a}\bbsp}_\bald{i}\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi].$ Theorem~\ref{repr} implies
${\bbsp\bbox{a}\bbsp}_\bald{i}=H^\#({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)$ for a function $H\in\cntm\xi.$ Let this
be forced by some $X\in\perfm$ such that ${\bbsp\bbox{x}\bbsp}\in X^\#$.
{\it We argue in ${\bbb{\bf M}\bbb}$.}
The set
${Y=\ans{x\in X:x(\bald{i})=H(x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)}}$ is a closed subset of $X$
because $H$ is continuous. Therefore by Corollary~\ref{bor}
there exists $X'\in\perf$ such that either $X'\subseteq Y$ or
$X'\subseteq X\setminus Y.$ The former possibility means that the
function $F(x)=x(\bald{i})$ is reducible to $\xi$ on $X'.$ Since $F$
is obviously reducible to $[\<\hspace{-2pt}\bald{i}],$ it is also reducible to
$\xi\cap[\<\hspace{-2pt}\bald{i}]$ by Lemma~\ref{inter}, therefore to
$[<\hspace{-2pt}\bald{i}],$ because $\bald{i}\not\in\xi.$ But this evidently
contradicts property~\ref{perf1} of $X'$.
This contradiction shows that in fact $X'\subseteq X\setminus Y.$ But
then $X'$ forces that ${\bbsp\bbox{x}\bbsp}(\bald{i})={\bbsp\bbox{a}\bbsp}_\bald{i}$ is {\it not\/} equal to
$H^\#({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi),$ contradiction with the choice of $X$ and
$X'$.\qed\vspace{3mm}
\noindent{\bf Proof of statement \ref{m2+} of Theorem~\ref{m}}\\[2mm]
We obtain the result from item~\ref{mm2+} of Theorem~\ref{mm}
using essentially the same type of reasoning as above.\vspace{3mm}
\noindent{\bf Proof of statement \ref{m3} of Theorem~\ref{m}}\\[2mm]
Let $\xi\in {\bbb{\bf M}\bbb}$ be an
initial segment of $\bI,$ and $c\in{\bbb{\bf N}\bbb}\cap{\skri D}.$ We have to prove
that either $c\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ or there exists $\bald{i}\not\in\xi$
such that ${\bbsp\bbox{a}\bbsp}_\bald{i}\in{\bbb{\bf M}\bbb}[c].$ Theorem~\ref{repr} tells that
$c=F^\#({\bbsp\bbox{x}\bbsp})$ for some $F\in\contm.$ Let this be forced by some
$X\in\perfm.$ We also assume that on the contrary $c$
{\it does not\/} satisfy requirement~\ref{m3} of
Theorem~\ref{m}, and this is forced by the same $X$.
{\it We argue in ${\bbb{\bf M}\bbb}$.}
It follows from Theorem~\ref{mm} (item~\ref{mm3}) that
there exists $X'\in\perf,$ $X'\subseteq X,$ such that either $F$ is
reducible to $\xi$ on $X'$ or $F$ captures some $\bald{i}\not\in\xi$
on $X'.$
Consider the ``either'' case. Then (see Remark~\ref{redr}) there
exists a function $H\in\cnt\xi$ such that $F(x)=H(x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)$ for
all $x\in X'.$ In this case $X'$ forces that $c\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi],$
contradiction with the choice of $X$ and $X'$.
Consider the ``or'' case. Then there exists a continuous function
$E:{\skri D}\;\lra\;{\skri D}$ such that $x(\bald{i})=E(F(x))$ for all $x\in X'.$
Then $X'$ forces ${\bbsp\bbox{a}\bbsp}_\bald{i}={\bbsp\bbox{x}\bbsp}(\bald{i})=E^\#(F^\#({\bbsp\bbox{x}\bbsp}))\in{\bbb{\bf M}\bbb}[F^\#({\bbsp\bbox{x}\bbsp})],$
again a contradiction with the choice of $X$ and $X'$.
{\it We argue in ${\bbb{\bf N}\bbb}$.}
It remains to note that the possibilities
of statement~\ref{m3} of Theorem~\ref{m} are incompatible by the
already proved statement~\ref{m2}.\qed
\subsubsection*{The Sacksness}
In this subsection we prove the principal item of Theorem~\ref{m}
--- statement~\ref{m1} which shows that in fact ${\bbb{\bf N}\bbb}$ is an
iterated Sacks extension of ${\bbb{\bf M}\bbb}$ with $\bI$ as the ``length'' of
the iteration.
\bte
\label{sack}
Every\/ ${\bbsp\bbox{a}\bbsp}_\bald{i}$ is Sacks generic over
${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}}]={\bbb{\bf M}\bbb}[\ang{{\bbsp\bbox{a}\bbsp}_\bald{j}:\bald{j}<\bald{i}}]$.
\ete
\noindent{\bft Proof\hspace{2mm} } Assume that $S\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}}]$ is, in
${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}}],$ a dense subset in the collection of all
perfect subsets of ${\skri D}=2^\om;$ we have to prove that
${\bbsp\bbox{a}\bbsp}_\bald{i}\in C^\#$ for some $C\in S.$ Suppose that on the contrary
some $X\in \perfm$ such that ${\bbsp\bbox{x}\bbsp}\in X^\#$ forces the opposite.
{\it We argue in ${\bbb{\bf M}\bbb}.$\/}
The set
$D(y)=D_{Xy}(\bald{i})=\ans{x(\bald{i}):x\in X\;\,\&\;\, x\rsd{<\bald{i}}=y}$ is perfect
for all $y\in Y=X\rsd{<\bald{i}}$ by property~\ref{perf1} of $X$.
{\it We argue in ${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}}].$\/}
Notice that
${\bbsp\bbox{y}\bbsp}={\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}}$ belongs to $Y^\#.$ Therefore $D^\#({\bbsp\bbox{y}\bbsp})$ is a
perfect set (certain absoluteness is applied). Thus there exists
a set $C\in S$ such that $C\subseteq D^\#({\bbsp\bbox{y}\bbsp})$.
Now we are in need of a coding of closed subsets of ${\skri D}=2^\om.$
Let $\ans{B_n:n\in\om}$ be a recursive enumeration of
all basic clopen sets in ${\skri D}.$ We put
$\clo{c}={\skri D}\setminus\bigcup_{c(n)=0}B_n$ for all $c\in 2^\om.$
Thus every closed $C\subseteq{\skri D}$ is equal to $\clo c$ for some
$c\in 2^\om$.
{\it We argue in ${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}}]$}.
We define $c\in 2^\om$ so
that $c(n)=0$ iff $B_n$ is a basic clopen set disjoint from $C;$
then $C=\clo{c}.$ By Theorem~\ref{repr}, $c=H^\#({\bbsp\bbox{y}\bbsp})$ for
some $H\in\cntm{<\bald{i}}.$ Since ${\bbsp\bbox{a}\bbsp}_\bald{i}={\bbsp\bbox{x}\bbsp}(\bald{i})\not\in C,$ we can
assume that $X$ forces that $\clo{H^\#({\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}})}$ is a perfect
subset of $D^\#({\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}})$ and ${\bbsp\bbox{x}\bbsp}(\bald{i})$ does not belong to
$\clo{H^\#({\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}})}$.
{\it We argue in ${\bbb{\bf M}\bbb}.$\/}
The continuation of the proof is
based on the following fact: the set
$$
U=\ans{x\in X: x(\bald{i})\in\clo{H(x\rsd{<\bald{i}})}}
$$
contains a subset
$U'\in\perf.$ Such a condition $U'$ would force that ${\bbsp\bbox{x}\bbsp}(\bald{i})={\bbsp\bbox{a}\bbsp}_\bald{i}$
belongs to $\clo{H^\#({\bbsp\bbox{x}\bbsp}\rsd{<\bald{i}})}$ by absoluteness,
contradiction with the statement forced by $X$.
Thus we concentrate on the mentioned key fact. Notice that
since we deal with compact spaces, the set
$$
Y_1=\ans{y\in Y:\clo{H(y)}\,\hbox{ is a perfect subset of }\,D(y)}
$$
is $\bbox{G}_\delta$ in $Y.$ Corollary~\ref{bor} says that there
exists a set $Y_2\in\pe{<\bald{i}},$ $Y_2\subseteq Y,$ such that either
$Y_2\subseteq Y_1$ or $Y_2\cap Y_1=\emptyset.$ Then
$X_2=X\cap (Y_2\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI)\in\perf$ by Lemma~\ref{apro}.
Suppose that $Y_2\cap Y_1=\emptyset.$ Then by absoluteness $X_2$
forces that $\clo{H^\#({\bbsp\bbox{x}\bbsp}\rsd{<i})}$ is not a perfect subset of
$D^\#({\bbsp\bbox{x}\bbsp}\rsd{<i}),$ contradiction with the statement forced by $X$.
Thus in fact $Y_2\subseteq Y_1.$ We have to restrict $Y_2$ a little bit
more. Notice that for any clopen $G\subseteq{\skri D}$ the set
$Y(G)=\ans{y\in Y_2:\clo{H(y)}\cap G\not=\emptyset}$ is closed.
Then by Corollary~\ref{prebor} there exists $Y'\subseteq Y_2,$
$Y'\in\pe{<\bald{i}}$ such that $Y'\cap Y(G)$ is clopen in $Y'$
for every $G$.
We demonstrate that
${Z=\ans{z\in X\rsd{\<\bald{i}}:z\rsd{<\bald{i}}\in Y'\;\,\&\;\, z(\bald{i})\in
\clo{H(z\rsd{<\bald{i}})}}}$
belongs to $\pe{\<\bald{i}}.$ To check all the requirements of
lemmas \ref{gath} and \ref{gath'} notice that $Z$ is closed and
$Z\rsd{<\bald{i}}=Y'\in\pe{<\bald{i}}.$ Moreover if $y\in Y'$ then
$D_{Zy}(\bald{i})=\clo{H(y)}$ is perfect since $Y'\subseteq Y_1,$ so that
$Z$ satisfies \ref{perf1}. Finally $Z$ also satisfies \ref{oz} by
the choice of $Y'.$ Therefore in fact $Z\in\pe{\<\bald{i}}$.
It remains to set $U'=X\cap (Z\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\bI)$.\qed\vspace{4mm}
This ends the proof of Proposition~\ref{counta} --- the countable
case in Theorem~\ref{m}.\qed
\newpage
\subsection{Uncountable case}
\label{uncount}
We carry out the general case of Theorem~\ref{m} in very brief
manner because the principal points can be reduced to the already
considered countable case.
Thus let ${\bbb{\bf M}\bbb}$ be a countable transitive model of $\ZFC,$
$\bI\in{\bbb{\bf M}\bbb}$ a po set.
Let $\cs$ be the collection of all sets $\xi\in{\bbb{\bf M}\bbb},$ $\xi\subseteq\bI,$
such that $\rbox{card}\,\xi\<\aleph_0$ in ${\bbb{\bf M}\bbb}.$ Notice that sets
$\xi\in\cs$ are, generally speaking, not initial segments of
$\bI$ or of each other.
For any $\xi\in\cs,$ let $\peM\xi=(\pe\xi)^{\bbb{\bf M}\bbb}.$ The set
$\perfm=\bigcup_{\xi\in\cs}\peM\xi$ is the forcing notion. To
define the order, we first put $\supp X=\xi$ whenever
$X\in\peM\xi.$ Now we set $X\<Y$ ($X$ is stronger than $Y$) iff
$\xi=\supp Y\subseteq \supp X$ and $X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\subseteq Y$.
Let $G\subseteq\perfm$ be a generic set over ${\bbb{\bf M}\bbb}.$ Then there exists
unique indexed set ${\bbsp\bbox{x}\bbsp}=\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\bI},$ all ${\bbsp\bbox{a}\bbsp}_\bald{i}$ belong
to ${\skri D},$ such that ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\in X^\#$ whenever $X\in G$ and
$\supp X=\xi.$ Moreover ${\bbb{\bf M}\bbb}[G]={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}]={\bbb{\bf M}\bbb}[\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\bI}]$.
\bpro
\label{m-m}
The model\/ ${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[G]={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}]$ satisfies Theorem~\ref{m}.
\epro
\noindent{\bft Proof\hspace{2mm} }{}is based on the two principal statements.
\bte
\label{gen1}
$\aleph_1^{\bbb{\bf M}\bbb}$ remains a cardinal in ${\bbb{\bf N}\bbb}$.~\footnote
{\rm\ The behaviour of other cardinals depends on the cardinal
structure in ${\bbb{\bf M}\bbb},$ the cardinality of $\bI,$ and the cardinality
of chains in $\bI.$ It is not our intension here to investigate
this matter.}
\ete
\bte
\label{gen2}
Assume that\/ $\bJ\in{\bbb{\bf M}\bbb}$ is an initial segment of\/ $\bI$ and\/
$c\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\bJ]\cap\om^\om.$ There exist\/
$\xi\in\cs,$ $\xi\subseteq\bJ,$ and a function\/ $H\in\cntm\xi$ such
that\/ $c=H^\#({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi)$.
\ete
Theorem~\ref{gen2} allows to repeat the reasoning in
Section \ref{re} and prove statements \ref{m1} through \ref{m3}
of Theorem~\ref{m} using properties of forcing conditions and
continuous functions proved mainly in Section~\ref{cont}.
Thus theorems \ref{gen1}
and \ref{gen2} suffice for Proposition~\ref{m-m} and
Theorem~\ref{m}.\qed\qed\vspace{3mm}
\noindent
{\bf Proof}\hspace{1mm} of Theorem~\ref{gen1}\\[2mm]
Let ${\underline f}$ be a name of a function defined on $\om$ in the
language of forcing. We fix $X_0\in\perfm.$ The aim is to
obtain a condition $X\in\perfm,$ $X\<X_0,$ and a countable in
${\bbb{\bf M}\bbb}$ set $R$ such that $X$ forces that the range of ${\underline f}$ is
included in $R.$ Let $\xi_0=\supp{X_0}$.\vspace{1mm}
{\it We argue un ${\bbb{\bf M}\bbb}$.}
To utilize the proof of Theorem~\ref{card} we reduce the forcing
of statements related to ${\underline f}$ to a certain $\zeta\in\cs.$
Let $\xi\in\cs.$ We say that a set ${\skri X}\subseteq\pe\zeta$ is {\it
adequate\/} if\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){\alph{enumi})}
\def\theenumi{\arabic{enumi})}
\item\label{a}
for any initial segment
$\eta
\ans{\bald{j}\in\xi: \bald{j}\not\>\bald{i}_1\;\,\&\;\,...\;\,\&\;\, \bald{j}\not\>\bald{i}_n},$
where $\bald{i}_1,...,\bald{i}_n\in\xi,$
and any pair $X,\,Y\in{\skri X},$ if $Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta\subseteq X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ then
$Z=X\cap (Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\eta)\in{\skri X},$ \hfill and\hfill\vspace{-1mm}
\item\label{b}
for all $X\in{\skri X},$ $\bald{i}\in\xi,$ $e\in\ans{0,1},$ the
set $X_e=\rbox{Spl}(X,\bald{i},e)$ belongs to ${\skri X}.$~\footnote
{\rm\ Notice that $Z\in\pe\xi$ and $X_e\in\pe\xi$ by
lemmas \ref{apro} and \ref{spl}.}\vspace{-1mm}
\end{enumerate}
It is obvious that every countable ${\skri X}'\subseteq\pe\xi$ can be extended
to a countable adequate ${\skri X}\subseteq\pe\xi$.
It can be easily verified that if $\xi\subseteq\zeta\in\cs$ and
$X\in\pe\xi$ then $X\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\zeta\in\pe\xi$ (although in general
Lemma~\ref{apro} is not true in the case when $\xi$ is not an
initial segment of $\zeta$). Therefore for all $\xi\in\cs,$
$n\in\om,$ and countable ${\skri X}'\subseteq\pe\xi$ there exists $\zeta\in\cs$
and an adequate countable ${\skri X}\subseteq\pe\zeta$ such that $\xi\subseteq\zeta$
and\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){\roman{enumi})}
\def\theenumi{\arabic{enumi})}
\item
$X'\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\zeta\in {\skri X}$ whenever $X'\in{\skri X}'$;\hfill and\hfill\vspace{-1mm}
\item
for any $X'\in{\skri X}'$ there exists $X\in{\skri X}$ such that $X\<X'$
and $X$ decides the value of ${\underline f}(n)$.\vspace{-1mm}
\end{enumerate}
This allows to start from $X_0\in\pe{\xi_0}$ and define by
induction a sequence $\xi_0\subseteq\xi_1\subseteq\xi_2\subseteq...$ of
$\xi_n\in\cs$ and a sequence of countable adequate
${\skri X}_n\subseteq\pe{\xi_n}$ such that\vspace{-1mm}
\begin{enumerate}
\def\arabic{enumi}){\arabic{enumi})}
\def\theenumi{\arabic{enumi})}
\item
$X\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\xi_{n+1}\in {\skri X}_{n+1}$ whenever $X\in{\skri X}_n$;\hfill
and\hfill\vspace{-1mm}
\item
for any $X'\in{\skri X}_n$ there exists $X\in{\skri X}_{n+1}$ such that
$X\<X'$ and $X$ decides the value of ${\underline f}(n)$.\vspace{-1mm}
\end{enumerate}
We set $\zeta=\bigcup_{n\in\om}\xi_n$ and
${\skri X}=\bigcup_{n\in\om}\ans{X\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\zeta:X\in{\skri X}_n}.$ Then
${\skri X}\subseteq\pe\zeta$ is a countable adequate family which satisfies the
property that
$$
\forall\,X'\in{\skri X}\;\forall\,n\;\exists\,X\in{\skri X}\;
[\,X\subseteq X'\;\,\&\;\, X\,\hbox{ decides the value of }\,{\underline f}(n)\,]\,.
$$
We notice now that the transformations of sets used in the proofs
of lemmas \ref{suz} and \ref{pand} are of types \ref{a} and
\ref{b}. Therefore arguing like in the proof of Theorem~\ref{card}
we can obtain a fusion sequence $\ang{X_u:u\in 2^{<\om}}$
of sets $X_u\in{\skri X}$ such that $X_\La\subseteq X_0$ and for all $m,$
every $X_u,\,\,u\in 2^m,$ decides the value of ${\underline f}(m).$ Then
$X=\bigcap_m\bigcup_{u\in 2^m}X_u\in\pe\zeta,$ $X\subseteq X_0,$ and $X$
forces that the range of ${\underline f}$ is a subset of a countable in
${\bbb{\bf M}\bbb}$ set.\qed\vspace{3mm}
\noindent
{\bf Proof}\hspace{1mm} of Theorem~\ref{gen1}\\[2mm]
Let ${\underline c}$ be a
name for $c$ containing an explicit absolute construction of $c$
from ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\bJ$ and some $p\in{\bbb{\bf M}\bbb}$.
{\it We argue in ${\bbb{\bf M}\bbb}.$}
Given $X_0\in\pe{\xi_0},$ we argue as in
the proof of Theorem~\ref{gen1} and get $\zeta\in\cs,$
$\xi_0\subseteq\zeta\subseteq\bJ,$ and a countable adequate ${\skri X}\subseteq\pe\zeta$
such that $X_0\hspace{-2pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}^{-1}\hspace{-1pt}\zeta\in{\skri X}$ and
$$
\forall\,X'\in{\skri X}\;\forall\,n\;\exists\,X\in{\skri X}\;
[\,X\subseteq X'\;\,\&\;\, X\,\hbox{ decides the value of }\,{\underline c}(n)\,]\,.
$$
It remains to carry out the construction in the proof of
Theorem~\ref{repr} within ${\skri X}$.\qed\vspace{3mm}
This ends the proof of Theorem~\ref{m} in general
case.\qed
\newpage
\newcommand{{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}}}{{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi}\hspace{1pt}}}
\newcommand{{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}}}{{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta}\hspace{1pt}}}
\subsection{The non--Glimm--Effros equivalence
relation}
\label{seqge}
This section is devoted entirely to the proof of Theorem~\ref{ge}.
Thus let $\mathbin{\relf{C}}$ be the equivalence relation defined on reals by
$x\mathbin{\relf{C}} y$ iff $\rbox{L}[x]=\rbox{L}[y]$.
We have to find a model for Theorem~\ref{ge}.
Let ${\bbb{\bf M}\bbb}$ be a countable transitive model of $\ZFC$ plus the
axiom of constructibility -- the initial model.
We define the ``length'' of the iteration $\bI=\om_1^{\bbb{\bf M}\bbb}\times\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}$
($\om_1^{\bbb{\bf M}\bbb}$ copies of $\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}},$ the integers). Thus from the point
of view of ${\bbb{\bf M}\bbb},$ $\bI$ is the set of all pairs $\ang{\al,z},$
$\al<\om_1$ and $z\in \mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}},$ linearly ordered lexicographically,
but of course not wellordered.
Let $\cs$ be the collection of all initial segments $\xi\in{\bbb{\bf M}\bbb},$
$\xi\subseteq\bI,$ such that $\rbox{card}\,\xi\<\aleph_0$ in ${\bbb{\bf M}\bbb}.$ (This
formally differs from the definition in Section~\ref{uncount},
but not essentially, since each \dd{\bbb{\bf M}\bbb} countable subset of $\bI$
can be covered by a countable initial segment.)
We define $\peM\xi=(\pe\xi)^{\bbb{\bf M}\bbb}$ (for all $\xi\in\cs$),
$\perfm=\bigcup_{\xi\in\cs}\peM\xi,$ the ``support'' $\supp X,$
and the order on $\perfm$ as in Section~\ref{uncount}.
Let us fix a generic over ${\bbb{\bf M}\bbb}$ set $G\subseteq\perfm.$ We define
${\bbsp\bbox{x}\bbsp}=\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\bI}$ (all ${\bbsp\bbox{a}\bbsp}_\bald{i}$ being elements of ${\skri D}$).
We prove that the model
${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[G]={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}]={\bbb{\bf M}\bbb}[\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\bI}]$ satisfies
Theorem~\ref{ge}.
\bte
\label{tge}
It is true in\/ ${\bbb{\bf N}\bbb}$ that\/ $\mathbin{\relf{C}}\;:$\vspace{-1mm}
\begin{itemize}
\item[--] neither admits a R-OD separating family$;$\vspace{-1mm}
\item[--] nor admits an uncountable R-OD pairwise\/
\dd\mathbin{\relf{C}} inequivalent set$.$
\end{itemize}
\ete
\noindent{\bft Proof\hspace{2mm} } Let us first investigate the structure of the degrees of
constructibility (i.e. \dd\mathbin{\relf{C}} degrees) in ${\bbb{\bf N}\bbb}$ -- or, that is the same
since ${\bbb{\bf M}\bbb}$ models $\rbox{V}=\rbox{L},$ degrees of \dd{\bbb{\bf M}\bbb} constructibility.
The set
$\bI$ has a nice property: its initial segments admit a clear description.
Indeed, each $\xi\in\cs$ is equal to one of the following:
$$
\begin{array}{rccccl}
\xi_{\al z} & = & [\<\hspace{-2pt}\bald{i}] & = &\ans{\bald{j}\in\bI:\bald{j}\<\bald{i}}\,, &
\hbox{where }\,\bald{i}=\ang{\al,z}\in\bI\,\hbox{ and }\,
\al<\om_1^{\bbb{\bf M}\bbb},\;\,z\in\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}\,;\\[2mm]
\xi_{\al} &&& = & \al\times\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}} & \hbox{for some }\,\al<\om_1^{\bbb{\bf M}\bbb}\,;
\end{array}
$$
obviously all of them belong to ${\bbb{\bf M}\bbb}.$ In particular, $\cs\in{\bbb{\bf M}\bbb}$ is
linearly ordered in ${\bbb{\bf M}\bbb}$ by inclusion. One can see that $\cs$ is order
isomorphic to $\om_1^{\bbb{\bf M}\bbb}\times(1+\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}})$.
\ble
\label{ndeg}
Suppose that\/ $c$ is a real in\/ ${\bbb{\bf N}\bbb}.$ There exists unique\/
$\xi\in\cs$ such that\/ ${\bbb{\bf M}\bbb}[c]={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$.
\ele
\noindent{\bft Proof\hspace{2mm} }{}of the lemma. (We recall that ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi=\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\xi}$.)
By Theorem~\ref{gen2}, $c=H^\#({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\zeta)$ for appropriate $\zeta\in\cs$
and $H\in\cnt\zeta$ in ${\bbb{\bf M}\bbb};$ in particular $c\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\zeta]$.
We set $\xi=\ans{\bald{i}\in\zeta:{\bbsp\bbox{a}\bbsp}_\bald{i}\in{\bbb{\bf M}\bbb}[c]};$ then $\xi\in\cs$ by
Theorem~\ref{m}, item~\ref{m1+}. (We mean: by Proposition~\ref{m-m}
which assures that ${\bbb{\bf N}\bbb}$ satisfies
Theorem~\ref{m}.) Furthermore items \ref{m2}, \ref{m2+},
\ref{m3} of the same theorem prove that both $c\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$
and ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi\in{\bbb{\bf M}\bbb}[c]$.\qed
\begin{corollary}\TF\
\label{cmany}
Suppose that\/ $c$ is a real in\/ ${\bbb{\bf N}\bbb}.$ There exists only
countably\/ {\rm(in ${\bbb{\bf N}\bbb}$)} many\/ \dd{\bbb{\bf M}\bbb} degrees below $c$.\qed
\end{corollary}
\subsubsection*{Proof of the ``nor'' part of Theorem~\ref{tge}}
Let, in ${\bbb{\bf N}\bbb},$ $S$ be a \dd\mathbin{\relf{C}} pairwise inequivalent subset of ${\skri D}$
defined (in ${\bbb{\bf N}\bbb}$) by a formula containing ordinals and a real
$p\in{\bbb{\bf N}\bbb}$ as parameters. By Lemma~\ref{ndeg} there exists an
initial segment, say $\eta=\xi_\al=\al\times\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}},$ $\al<\om_1^{\bbb{\bf M}\bbb},$
such that $p\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta].$ Then $S$ is definable in ${\bbb{\bf N}\bbb}$
using a formula containing ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta=\ang{{\bbsp\bbox{a}\bbsp}_\bald{i}:\bald{i}\in\eta}$ and
ordinals as parameters.\vspace{-1mm}
\begin{itemize}
\item[$(1)$] {\it We assert that\/ $S\subseteq{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta]$}.\vspace{-1mm}
\end{itemize}
This assertion implies that $S$ is countable by
Corollary~\ref{cmany}, therefore suffices for the ``nor'' part of
the theorem.
To prove the assertion, let us fix a real $r\in S$ in ${\bbb{\bf N}\bbb}.$ Then
we have $\rbox{L}[r]=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ for certain (unique) $\xi\in\cs$ by
Lemma~\ref{ndeg}.
Let $\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,k,l)$ be the formula:\vspace{-1mm}
\begin{itemize}
\item there exists a real $r'\in S$ such that
$\rbox{L}[r']=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ and $r'(k)=l$.\vspace{-1mm}
\end{itemize}
(${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ enters the formula via a definition of $S.$ We recall
that $\rbox{V}=\rbox{L}$ is assumed in ${\bbb{\bf M}\bbb},$ so that $\rbox{L}[...]$ in ${\bbb{\bf N}\bbb}$
corresponds to ${\bbb{\bf M}\bbb}[...]$ in the universe.) Then,
in ${\bbb{\bf N}\bbb},$ we have $r(k)=l$ iff $\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,k,l)$ for
all $k,\,l$.
Let ${\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}}$ and ${\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}}$ be the names for ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$ and
${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$.\vspace{-1mm}
\begin{itemize}
\item[$(2)$] {\it We assert that, for all\/ $k,\,l\in\om$ and\/
$X\in\perf,$ if $\eta\subseteq\supp X$ and\/ $X$ forces\/
$\varphi({\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}},{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}},k,l)$ then\/ $X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ already forces
$\varphi({\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}},{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}},k,l)$.}\vspace{-1mm}
\end{itemize}
One can easily see that $(2)$ implies $r\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta],$ that
is, implies $(1);$ therefore we concentrate on the assertion $(2)$.
Assume that $(2)$ is not true. Thus results in a pair of conditions
$X,\,Y\in \peM\zeta,$ where $\zeta\in\cs,$ $\eta\subseteq\zeta,$ such that
$X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta=Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,$ $X$ forces $\varphi({\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}},{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}},k,l),$ but $Y$
forces the negation of $\varphi({\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}},{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}},k,l),$ for some $k,\,l$.
In ${\bbb{\bf M}\bbb},$ both $X$ and $Y$ are members of $\pe\zeta.$ Let $F\in{\bbb{\bf M}\bbb}$
be a homeomorphism $X$ onto $Y$ satisfying requirements \ref{h1},
\ref{h2}, \ref{h3} of Theorem~\ref{thom}, in particular,
$x{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta=F(x){\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ for all\/ $x\in X,$ because
$X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta=Y{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$.
(Let us forget temporarily that a generic set $G\subseteq\perfm$ was fixed
above.) The homeomorphism $F$ induces the total automorphism of
the part of $\perfm$ stronger than $X$ onto
the part of $\perfm$ stronger than $Y,$ which results in a pair of
\dd\perfm generic over ${\bbb{\bf M}\bbb}$ sets $G,\,G'\subseteq\perfm$ and corresponding
${\bbsp\bbox{x}\bbsp},\,{\bbsp\bbox{x}\bbsp}'\in\can{\bI}$ such that $X\in G,$ $Y\in G',$ ${\bbb{\bf M}\bbb}[G]={\bbb{\bf M}\bbb}[G'],$
${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta={\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,$ and finally
${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ for all $\xi\in\cs.$ Thus we have
got one and the same generic extension ${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[G]={\bbb{\bf M}\bbb}[G']$
using two different generic sets.
Notice that the statement $\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,k,l)$ is true
while $\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,k,l)$ is false in ${\bbb{\bf N}\bbb}$ by the
choice of $X,\,Y.$ We cannot assert that ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi={\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$
(unless $\xi\subseteq\eta,$ of course), but the formula $\varphi$ was
defined so that it is \dd\mathbin{\relf{C}} invariant on the argument ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi:$
in other words,
$$
\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,k,l)\;\;\llra\;\;
\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,k,l)
$$
provided ${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi].$ Since this assumption
was obtained above, we conclude that
$\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,k,l)$ must be true in ${\bbb{\bf N}\bbb},$
contradiction.
This ends the proof of the ``nor'' part of Theorem~\ref{tge}.
\subsubsection*{Proof of the ``neither'' part of
Theorem~\ref{tge}}
Suppose that, on the contrary, $\mathbin{\relf{C}}$ admits in ${\bbb{\bf N}\bbb}$ a R-OD
separating family $\ang{X_\al:\al<\gamma},$ $\gamma$ an ordinal. As
above, then the family is definable by a formula containing
ordinals and some ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,$ $\eta\in\cs,$ as parameters. We
define, in ${\bbb{\bf N}\bbb}$,
$$
U(r)=\ans{\al<\gamma:r\in X_\al}
$$
for each real $r\in{\bbb{\bf N}\bbb}.$ Thus, $x\mathbin{\relf{C}} y$ iff $U(x)=U(y)$ for each pair
of reals $x,\,y$ in ${\bbb{\bf N}\bbb}$.\vspace{-1mm}
\begin{itemize}
\item[$(3)$] {\it We assert that\/ $U(r)\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta]$ for
all reals $r$ in ${\bbb{\bf N}\bbb}$}.\vspace{-1mm}
\end{itemize}
Generally speaking, one would expect that $U(r)$ needs $r,$ or
at least the \dd\mathbin{\relf{C}} degree of $r$ as a parameter of definition.
However, the \dd\mathbin{\relf{C}} degrees in ${\bbb{\bf N}\bbb}$ form a quite regular structure
by Lemma~\ref{ndeg}, so that
each degree is ``almost'' ordinal definable (but actually
{\it not\/} OD), which makes it possible to prove $(3)$.
As before, in the proof of the ``nor'' part, we reduce $(3)$ to a
forcing assertion. Let us fix a real $r\in{\bbb{\bf N}\bbb};$ then by
Lemma~\ref{ndeg} there exists $\xi\in\cs$ such that
$\rbox{L}[r]=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi].$ Let $\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,\al)$
be the formula:\vspace{-1mm}
\begin{itemize}
\item there exists a real $r'$ such that
$\rbox{L}[r']=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ and $r'\in X_\al$.\vspace{-1mm}
\end{itemize}
(${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ enters the formula via the enumeration of sets
$X_\al$.) Then, in ${\bbb{\bf N}\bbb},$ we have $\al\in U(r)$ iff
$\varphi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,\al)$ for all $\al$.\vspace{-1mm}
\begin{itemize}
\item[$(4)$] {\it We assert that, for all\/ $\al<\gamma$ and\/
$X\in\perf,$ if $\eta\subseteq\supp X$ and\/ $X$ forces\/
$\varphi({\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}},{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}},\al)$ then\/ $X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ already forces
$\varphi({\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\eta}\hspace{1pt}},{\hspace{1pt}\underline{{\bbsp\bbox{x}\bbsp}\res\xi}\hspace{1pt}},\al)$.}\vspace{-1mm}
\end{itemize}
As in the proof of the ``nor'' part above, $(4)$ implies
$U(r)\in{\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta],$ that
is, implies $(3);$ therefore it suffices to prove $(4).$ We omit the
reasoning because it is a copy of the proof of $(2)$ above: the
principal point is that the formula $\varphi$ is again \dd\mathbin{\relf{C}} invariant
on the argument ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi$.
Thus we obtain $(4)$ and $(3)$.
We continue the proof of the ``either'' part. The key moment is as
follows. It follows from assertion $(3)$ that in ${\bbb{\bf N}\bbb}$ each \dd\mathbin{\relf{C}}
degree is definable by a formula using only ordinals and ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$
as parameters. In particular, ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ plus ordinals as parameters
is enough to distinguish all \dd\mathbin{\relf{C}} degrees from each other.
This will help us to engineer a contradiction. The special mechanism
of getting a contradiction is based on the existence of order
automorphisms (shiftings) in each \dd\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}} group in
$\bI=\om_1^{\bbb{\bf M}\bbb}\times\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}$.
There exists an ordinal $\al<\om_1^{\bbb{\bf M}\bbb}$ such that
$\eta\subseteq\xi_\al=\al\times\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}.$ We set $\bald{i}=\ang{\al,0},$
$\bald{i}'=\ang{\al,1}$ -- two neighbouring elements in the least \dd\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}
group not participating in $\xi_\al.$ We set $\xi=[\<\hspace{-2pt}\bald{i}],$
$\xi'=[\<\hspace{-2pt}\bald{i}'].$ Since $\bald{i}'\in\xi'\setminus\xi,$ we have
${\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]\not={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'].$ (Item \ref{m2} of
Theorem~\ref{m} via Proposition~\ref{m-m}.) Take a pair of reals
$r,\,r'\in{\bbb{\bf N}\bbb}$ such that ${\bbb{\bf M}\bbb}[r]={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ and
${\bbb{\bf M}\bbb}[r']={\bbb{\bf M}\bbb}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'];$ then $\rbox{L}[r]\not=\rbox{L}[r']$ in ${\bbb{\bf N}\bbb},$ hence
$U(r)\not=U(r').$ Since both $U(r)$ and $U(r')$ belong to
$\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta]$ in ${\bbb{\bf N}\bbb}$ by $(3),$ we conclude that there exists a
formula $\psi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,x)$ containing only ordinals and
${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ as parameters, and such that the following is true
in ${\bbb{\bf N}\bbb}$ for every real $x$:
$$
\rbox{L}[x]=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]\;\lra\;\psi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,x))
\hspace{8mm}\hbox{and}\hspace{8mm}
\rbox{L}[x]=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi']\;\lra\;\neg\;\psi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,x))\,.
\eqno{(\ast)}
$$
Therefore, a condition $X\in G$ forces $(\ast).$ One can w.l.o.g.
assume that $\eta\subseteq \supp X.$\vspace{-1mm}
\begin{itemize}
\item[$(5)$] {\it We assert that the weaker condition
$Y=X{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta$ forces $(\ast)$.}\vspace{-1mm}
\end{itemize}
This is an assrtion of the same type as $(2)$ and $(4)$ above; its
proof does not differ from the proof of $(2)$.
Notice that $Y\in G$.
Let us consider the order automorphism $h:\bI\,\hbox{ onto }\,\bI$
defined as follows: $h(\ang{\al,k})=\ang{\al,k+1}$ for the given
$\al$ and each $k\in\om,$ and $h(\ang{\beta,k})=\ang{\beta,k}$
whenever $\beta\not=\al.$ (Then $h(\bald{i})=\bald{i}'$.) Thus $h$ shifts only
the \dd\al th copy of $\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}$ in $\bI$ but does not move anything else.
The $h$ generates an order automorphism
$Z\,\longmapsto\,Z':\perfm\,\hbox{ onto }\,\perfm$ in obvious way.
We observe that $Y'=Y$ because $\supp Y=\eta\subseteq\xi_\al=\al\times\mathord{{\sf Z}\hspace{-4.5pt}{\sf Z}}$.
We set $G'=\ans{Z':Z\in G}.$ Then $Y\in G',$ $G'$ is \dd\perfm
generic over ${\bbb{\bf M}\bbb},$ and moreover, ${\bbb{\bf M}\bbb}[G']={\bbb{\bf M}\bbb}[G]$ because $h\in{\bbb{\bf M}\bbb}.$
Let ${\bbsp\bbox{x}\bbsp}'=\ang{{\bbsp\bbox{a}\bbsp}'_\bald{j}:\bald{j}\in\bI}$ be defined from $G'$ as ${\bbsp\bbox{x}\bbsp}$ was
defined from $G.$ Then ${\bbsp\bbox{a}\bbsp}'_\bald{j}={\bbsp\bbox{a}\bbsp}_{h(\bald{j})}$ for all $\bald{j};$ in
particular \ (a) ${\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta={\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta\,,$ \ and \ (b)
${\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi'$ is a shift of ${\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi,$ so that
$\rbox{L}[{\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi']=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ in ${\bbb{\bf N}\bbb}={\bbb{\bf M}\bbb}[G]={\bbb{\bf M}\bbb}[G']$.
It follows from (a) and the choice of $Y$ that
$\neg\;\psi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,x)$ holds in ${\bbb{\bf N}\bbb}$ provided a real $x$
satisfies $\rbox{L}[x]=\rbox{L}[{\bbsp\bbox{x}\bbsp}'{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi']$ in ${\bbb{\bf N}\bbb}.$ On the other hand, we
have already got $\psi({\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\eta,x)$ in ${\bbb{\bf N}\bbb}$ provided
$\rbox{L}[x]=\rbox{L}[{\bbsp\bbox{x}\bbsp}{\hspace{1.5pt}\mathbin{\hspace{0.1ex}|\hspace{0.1ex}}\hspace{1.5pt}}\xi]$ holds in ${\bbb{\bf N}\bbb}.$ These two statements
contradict each other by (b).
This ends the proof of the ``neither'' part of
Theorem~\ref{tge}.\qed\vspace{4mm}
This also ends the proof of Theorem~\ref{ge}.\qed
|
\section{INTRODUCTION}
\subsection{Motivation}
A wide range of quasi-one-dimensional materials undergo a
structural transition, known as the Peierls or charge-density-wave
(CDW) transition, as the temperature is lowered \cite{gru,con,gor,car}.
A periodic lattice distortion, with wave vector, $2k_F$, twice that of
the Fermi wavevector, develops along the chains.
Anomalies are seen in the electronic properties
due to the opening of an energy gap over the Fermi surface.
Over the past decade, due to the development of high-quality
samples and higher resolution experimental techniques,
new data has become available
which allows a quantitative comparison of experiment with theory.
The most widely studied material is the blue bronze, K$_{0.3}$MoO$_3$.
There is a well-defined three-dimensional transition at
$T_P=183$ K and careful measurements have been made of
thermodynamic anomalies \cite{bri} and CDW coherence lengths \cite{gir}
at the transition. The critical region, estimated from the
Ginzburg criterion \cite{gin} is only a few percent of
the transition temperature and so the transition should be described
by an anisotropic three-dimensional Ginzburg-Landau free energy
functional, except close to the transition temperature.
The challenge is to derive from a microscopic theory
the coefficients in the Ginzburg-Landau free energy
so a quantitative comparison
can be made between theory and experiment.
Inspiration is provided by the case of superconductivity.
The superconducting transition is well described by
Ginzburg-Landau theory and the coefficients can be calculated
from BCS theory \cite{schr} and depend on microscopic
parameters such as the normal state density of states,
Debye frequency, and the electron-phonon coupling.
This program is so successful that one can even consider
refinements to BCS theory, such as strong coupling effects,
in order to get better agreement between experiment and
theory \cite{carb}.
However, the problem of the CDW transition is more difficult
because of the large fluctuations due to the quasi-one dimensonality.
\subsection{Ginzburg-Landau theory}
The Peierls transition is described by an order parameter
which is proportional to the $2k_F$
lattice distortion along the chains.
The order parameter is complex if the lattice distortion
is incommensurate with the lattice. For a commensurate
lattice distortion (e.g., a half-filled band) the order
parameter is real.
I recently considered the general problem of Ginzburg-Landau
theory for a three-dimensional phase transition,
described by a complex order parameter,
in a system of weakly coupled chains \cite{mck2}.
The key results of that study are now summarized,
partly to put this paper in a broader context.
The Ginzburg-Landau free energy functional $F_1[\phi]$
for a {\it single} chain with a complex order
parameter $\phi(z)$, where $z$ is the co-ordinate along
the chain, is
\begin{equation}
F_1[\phi]=\int dz \left[
a \mid\phi\mid^2 + \ b \mid\phi\mid^4 +
\ c \mid {\partial \phi\over \partial z}\mid ^2 \right].
\label{aa1}
\end{equation}
Near the single chain mean-field transition temperature
$T_0$ the second-order coefficient $a(T)$ can be written
\begin{equation}
a(T)= a^\prime \left( {T \over T_0} - 1 \right).
\label{aa10}
\end{equation}
Due to fluctuations in the order parameter
this one-dimensional system cannot develop long-range order at
finite temperature \cite{lan,sca}.
To describe a finite-temperature
phase transition, consider a set of weakly interacting
chains. If $\phi_i(z)$ is the order parameter on the $i$-th chain
the free energy functional for the system is
\begin{equation}
F[\phi_i(z)]=\sum_i F_1[\phi_i(z)] -
{J \over 4} \sum_{<i,j>} \int dz {\rm Re} [\phi_i(z)^* \phi_j(z)]
\label{ad1}
\end{equation}
where $J$ describes the interchain interactions
between nearest neighbours.
A mean-field treatment of this functional will only
give accurate results if the width of the three-dimensional
critical region is much smaller than $T_0$. This requires
that the width of the one-dimensional critical region $\Delta t_{1D}
\equiv (bT_0)^{2/3}/a^\prime c^{1/3}$
be sufficiently small that
\begin{equation}
\Delta t_{1D} \ll \left( { J \over a^\prime} \right)^{2/3}.
\label{ad10}
\end{equation}
If this is not the case one can integrate out
the one-dimensional fluctuations to derive a new Ginzburg-Landau
functional with renormalized coefficients,
\begin{equation}
\tilde F[\Phi(x,y,z)]= {1 \over a_x a_y}\int d^3 x \left[
A \mid \Phi \mid^2
+B \mid \Phi \mid^4
+ C_x \mid {\partial \Phi \over \partial x }\mid^2
+C_y \mid {\partial \Phi \over \partial y }\mid^2
+ C_z \mid {\partial \Phi \over \partial z }\mid^2
\right]
\label{bg1}
\end{equation}
where $a_x$ and $a_y$ are the lattice constants
perpendicular to the chains.
The new order parameter $\Phi(x,y,z)$, is proportional
to the average of
$\phi_i(z)$ over neighbouring chains.
The three-dimensional mean-field temperature $T_{3D}$ is defined
as the temperature at which the the coefficient $A(T)$
changes sign.
Close to $T_{3D}$
\begin{equation}
A=A^\prime\left({T \over T_{3D}} -1 \right).
\label{abg1}
\end{equation}
The transition temperature $T_{3D}$ and the
coefficients $A^\prime$, $B$, $C_x$, $C_y$, and $C_z$
can be written in terms of the interchain interaction
$J$ and the coefficients
$a$, $b$, and $c$ of a single chain.
The coefficients in (\ref{bg1}) determine measurable quantities associated
with the transition such as the specific heat jump,
coherence lengths and width of the critical region.
Most of the physics is
determined by a {\it single} dimensionless parameter
\begin{equation}
\kappa \equiv { 2 (bT)^2 \over |a|^3 c}.
\label{aat1}
\end{equation}
which is a measure of the fluctuations along a single chain.
It was assumed that the coefficients $a$, $b$, and $c$
were independent of temperature and the measurable
quantities at the transition were determined as a function
of the interchain coupling. The transition temperature
increases as the interchain coupling increases. The coherence
length and specific heat jump depends only on the
single chain coherence length, $\xi_0 \equiv (c /|a|)^{1/2}$,
and the interchain coupling.
The width of the critical region, estimated from the
Ginzburg criterion, was virtually parameter independent,
being about 5-8 per cent of the transition temperature for
a tetragonal crystal. Such a narrow critical region is
consistent with experiment, and shows that Ginzburg-Landau
theory should be valid over a broad temperature range.
This paper uses a simple model to demonstrate
some of the difficulties involved in deriving the coefficients
$a$, $b$, $c$, and $J$ from a realistic microscopic theory.
\subsection{Microscopic theory}
The basic physics of quasi-one-dimensional CDW
materials is believed to be described by a Hamiltonian due
to Fr\"ohlich \cite{fro} which describes electrons
with a linear coupling to phonons.
Even in one dimension this is a highly non-trivial
many-body system and must treated by some approximation scheme.
The simplest treatment \cite{fro,ric0,all} is a rigid-lattice one
in which the phonons associated with the lattice
distortion are treated in the mean-field approximation
and the zero-point and thermal lattice motions are neglected.
The resulting
theory is mathematically identical to BCS theory \cite{all}.
An energy gap opens
at the Fermi surface at a temperature
$T_{RL} \simeq 1.14E_F e^{-1/ \lambda}$ where $E_F$ is
the Fermi energy and $\lambda$ is the dimensionless
electron-phonon coupling.
$T_{RL}$ is related to the
zero-temperature energy gap $\Delta_{RL}(0)$ by
\begin{equation}
\Delta_{RL}(0) = 1.76 k_B T_{RL}.
\label{trl}
\end{equation}
In this approximation the coefficients in the single-chain
Ginzburg -- Landau free energy functional (\ref{aa1}) are \cite{all}
\begin{equation}
a_{RL}(T)= {1 \over \pi v_F} \ln \left({T \over T_{RL}}\right)
\label{mfa}
\end{equation}
\begin{equation}
b_{RL}(T)= {1 \over \pi v_F} {7\zeta(3)\over(4\pi T)^2}
\label{mfb}
\end{equation}
\begin{equation}
c_{RL}(T)={1 \over \pi v_F}{7\zeta(3)v_F^2\over(4\pi T)^2}
\label{mfc}
\end{equation}
where $v_F$ is the Fermi velocity and $\zeta(3)$ is
the Riemann zeta function.
If $4t_\perp$ is the electronic bandwidth perpendicular to the
chains (see (\ref{fk}))
then the interchain coupling is given by \cite{sch,hor}
\begin{equation}
J_{RL}= \left( {4 t_\perp \over v_F}\right)^2 c_{RL}(T).
\label{mfc2}
\end{equation}
It might be hoped that the transition in real materials
can be described by the mean-field theory of the
functional (\ref{aa1}) with the coefficients (\ref{mfa}-\ref{mfc}).
However, this is not the case for several reasons.
(i) The width of the critical regime given by the one-dimensional
Ginzburg criterion \cite{ma} is very large:
$\Delta t_{1D} = 0.8 $ \cite{sch}, suggesting that
fluctuations are important because condition (\ref{ad10}) is
not satisfied.
(ii) A rigid-lattice treatment predicts a metallic density of
states at all temperatures above $T_{RL}$.
In contrast, magnetic susceptibility \cite{sco,joh,joh3},
optical conductivity \cite{deg,deg2,dre,dre2,bru,ber},
and photoemission \cite{dar,dar2,hwu} measurements suggest
that there is a gap
or pseudogap in the density of states for a broad temperature range
above $T_P$.
(iii) The transition temperature, specific heat jump, and coherence
lengths are inconsistent with rigid lattice predictions
(Table \ref{table1}).
This failure should not be surprising given that recent work has
shown that in the three-dimensionally ordered Peierls state
the zero-point and thermal lattice motions must
be taken into account to obtain a quantitative
description of the optical properties \cite{deg,deg2,mck,kim,lon}.
The next level of approximation is to use the coefficients
(\ref{mfa}-\ref{mfb}) and take into account the intrachain
order parameter fluctuations and the interchain coupling
and use results similar to those in References \cite{mck2}.
This is the approach that has been taken previously
\cite{sch,lee,die}.
There are two problems with this approach. First, if
the dimensionless parameter $\kappa$, given by (\ref{aat1}), is evaluated
using the expressions (\ref{mfa}-\ref{mfc})
the result is
\begin{equation}
\kappa_{RL}(T)= { 7 \zeta(3) \over 8 |\ln (T/T_{RL})|^3 }.
\label{at1}
\end{equation}
Hence, the temperature dependence is quite different
from the dependence
$\kappa \sim T^2$ that was assumed in References \cite{mck2,die,sca2}
and the analysis there needs to be modified.
The second and more serious problem is
one of self-consistency. The coefficients $a$, $b$, and $c$
are calculated neglecting fluctuations in the order
parameter which will modify the electronic properties
which in turn will modify the coefficients.
In this paper a simple model is used to demonstrate that
the fluctuations have a significant effect on the
single-chain coefficients.
An alternative microscopic theory, due to Schulz \cite{sch},
and which takes into account fluctuations in only the phase
of the order parameter
is briefly reviewed in Appendix \ref{appsch}.
\subsection{Overview}
Discrepancies between phonon rigid-lattice
theory and the observed properties of the Peierls state well
below the transition temperature $T_P$
were recently resolved \cite{mck,kim} by taking into account the
effect of the zero-point and thermal lattice motion on
the electronic properties. It was shown that the lattice fluctuations
have an effect similar to a Gaussian random potential. This mapping
breaks down near the transition temperature because of the phonon dispersion
due to the softening of the phonons near $2k_F$. In this paper
this dispersion is taken into account and the effect of the
large thermal lattice motion near the transition temperature
is studied.
The thermal lattice motion has the same effect on the electronic properties
as a static random potential with finite correlation length.
Close to the transition temperature, the problem reduces to
a simple model, corresponding to a single classical phonon,
which can be treated exactly (Section \ref{secham}).
This model was first studied by Sadovsk\~i\~i \cite{sad}.
It was recently used in the description of the destruction of
spin-density-wave states by high magnetic fields \cite{mck0}.
The one-electron Green's function is calculated in Section \ref{secgreen}.
There is a pseudogap in the density of states
(Fig \ref{figdos}).
The complexity of this simple model is indicated by two non-trivial
many-body effects: (i) Perturbation theory diverges
but is Borel summable. (ii) The traditional quasi-particle
picture breaks down
(Figure \ref{figspec}),
reminiscent of behaviour seen in Luttinger liquids\cite{voi}.
To illustrate that calculations based on perturbation
theory can be unreliable it is shown that a predicted scaling
relation between the specific heat and the temperature
derivative of the magnetic susceptibility \cite{cha} does not hold
if the {\it exact}, rather than approximate, density of
states is used in the calculation.
Using this model the coefficients $a$, $b$, and $c$
are calculated in Appendix \ref{seccoeff}.
The coefficients deviate significantly from the
rigid-lattice values if the pseudogap is
comparable to or larger than the transition temperature.
In Section \ref{secest} experimental data
is used to estimate the pseudogap in
K$_{0.3}$MoO$_3$.
\subsection{Previous work on fluctuations and the pseudogap}
To put this paper in context some important earlier work
is briefly reviewed.
Lee, Rice and Anderson \cite{lee}
considered how fluctuations in the order parameter
produce a pseudogap in the density of states.
It is important to be aware of the assumptions
made in their calculation. Although their results describe
much of the physics on a qualitative level, for the
reasons described below, their results cannot be expected to give
a quantitative description of the density of states near
the CDW transition.
The starting point of Lee, Rice, and Anderson
was the one-dimensional
Ginzburg-Landau functional (\ref{aa1}) with a {\it real}
order parameter and with the coefficients derived from
rigid-lattice theory (see equations (\ref{mfa}-\ref{mfc})).
Earlier, Scalapino, Sears, and Ferrell \cite{sca}
evaluated the correlation length $\xi_\parallel(T)$
for one-dimensional Ginzburg-Landau theory
with an exact treatment of the fluctuations in the
order parameter;
$\xi_\parallel(T)$ only diverges as $T \to 0$.
The results of this calculation were used by Lee, Rice, and Anderson as
input in a random potential
with correlations given by
\begin{equation}
<\Delta(z)\Delta(z')>=\Delta_{RL}
(T)^{2} \exp(-\mid z-z^{'}\mid /\xi_\parallel(T)) \label{af}
\end{equation}
where $\Delta_{RL}(T)$ is the rigid-lattice (BCS) order parameter
and the average is over the thermal fluctuations of the
order parameter.
The electronic
Green's function was calculated using equation (\ref{af}) and a
formula originally used for liquid
metals (essentially, second-order perturbation theory
for the random potential).
They found a gradual appearance of a gap as the
temperature decreased. For $T_P < 0.25 T_{RL},$ an absolute gap of
magnitude $\Delta_{RL}(0)$ appears.
Lee, Rice and Anderson
suggested that a three-dimensonal transition
occurs for $T_P \simeq 0.25 T_{RL}$ based on the
temperature at which $\xi_\parallel(T)$ becomes extremely large.
There are several problems with trying to use these results
to give a quantitative description of the CDW transtion,
because of the following assumptions.
(i) {\it A real order parameter.}
Most CDW transitions are described by a complex order parameter,
for which quantitatively distinct behaviour occurs.
For example, the transition to very large correlation
lengths for $T_P \simeq 0.25 T_{RL}$ does not occur
for a complex order parameter. (See Figure 6 in Reference
\cite{sca}).
(ii) {\it Rigid-lattice coefficients.}
It is shown in this paper that the pseudogap due to the
thermal lattice motion causes the Ginzburg-Landau coefficients to deviate
significantly from their rigid-lattice values (Figure \ref{figcoeff}).
(iii) {\it Perturbation theory.}
It is demonstrated in this paper that this is unreliable.
In particular as $\xi_\parallel(T) \to \infty $ in (\ref{af})
only a pseudogap rather than an absolute gap
develops in the density of states (Figure \ref{figdos}).
Rice and Str\"assler \cite {ric} calculated the contribution of the phonon
fluctuations to the electronic self energy in the Migdal approximation,
i.e., second-order perturbation theory.
Interchain interactions were included through an anisotropic phonon
dispersion. They found
a pseudogap in the density of states
above the transition temperature. At $T_P$ there is an absolute gap
whose magnitude is determined by the electron-phonon coupling
and the interchain interactions.
They equated the observed transition temperature
with the single-chain mean-field transition temperature $T_0$
which they found to be significantly
reduced below the rigid-lattice value $T_{RL}$
and to vanish as the interchain coupling vanishes.
In the limit of weak interchain interactions the analytic
form of the density of states
is identical to that of Lee, Rice, and Anderson \cite{lee}.
However, it is not commonly appreciated that the
origin of the pseudogap in the two calculations is quite different.
The magnitude of the Rice and Str\"assler pseudogap
is proportional to the thermal lattice motion (compare Section \ref{sectherm}),
while the pseudogap studied by Lee, Rice, and Anderson pseudogap is
by assumption equal to the rigid-lattice gap $\Delta_{RL}(T)$.
Calculations similar to that of Rice and Str\"assler have been
performed by Bjeli\~s and Bari\~si\~c \cite {bje}, Suzumura and Kurihara
\cite {suz}, Patton and Sham \cite {sha}, and Chandra \cite{cha}.
The main problem with these calculations are that they are
based on perturbation theory.
\section {MODEL HAMILTONIAN}
\label{secham}
The starting point for this paper is the
following one-dimensional model. The states in
an electron gas with Fermi velocity $v_F$
are described by spinors $\Psi(z)$. The upper and
lower components describe left and right moving electrons,
respectively.
The phonons are described by the field
\begin{equation}
\Delta(z) = g \sum_q \sqrt {\hbar \over 2M \omega_{2k_F+q}}
(b_{2k_F+q} + b_{-2k_F-q}^\dagger ) e^{iqz}
\end{equation}
where $b_s$ destroys a phonon of momentum $s$ and frequency $\omega_s$
and $g$ is the linear electron-phonon coupling.
The dimensionless electron-phonon coupling $\lambda$ is defined by
\begin{equation}
\lambda = 2 g^2 a_z/\pi v_F \omega_Q \label {bd}
\end{equation}
where $a_z$ is the lattice constant along the chains.
The electronic part of the Hamiltonian is \cite{bra}
\begin{equation}
H_{el} = \int dz \Psi^\dagger (z) \bigg[ - iv_F \sigma_3
{\partial \over \partial z} + {1 \over 2}(\Delta(z) \sigma_+ + \Delta(z)^*
\sigma_-)\bigg] \Psi(z)
\label{hamel}
\end{equation}
where $\sigma_3$ and
$ \sigma_{\pm} \equiv \sigma_1 \pm i \sigma_2$ are Pauli matrices.
This paper focuses on the following model where $\Delta(z)$
is replaced with a random potential with zero mean
and finite length correlations
\begin{equation}
\langle \Delta(z)\rangle = 0 \ \ \ \; \ \ \ \ \
\langle \Delta(z)\Delta(z')^* \rangle = \psi^2
\exp(-|z-z'|/\xi_\parallel).
\label{cor2}
\end{equation}
$\xi_\parallel$ is the CDW correlation length along the chains.
In most of this paper $\psi$ will be treated as a parameter.
It is central to this paper, being
a measure of the thermal lattice motion and
a measure of the pseudogap in the density of states.
This paper focuses on behaviour near $T_P$ and so the
limit $\xi_\parallel \psi/v_F \to \infty$ is taken.
A rough argument is now given to justify using this
model to describe thermal lattice motion near the phase transition.
\subsection {Thermal lattice motion}
\label{sectherm}
In rigid-lattice theory $\Delta(z)$ is replaced by its expectation value
$\langle \Delta(z) \rangle =\Delta_o$.
To go beyond this
the effect of the quantum and thermal
lattice fluctuations in the Peierls state was
recently modelled \cite{mck,kim,mck1} by treating $\Delta(z)$
as a static random potential with mean $\Delta_o$ and correlations
\begin{equation}
\langle \Delta(z)\Delta(z')^* \rangle =\Delta_o^2 + \gamma \delta(z-z')
\label{corprl}
\end{equation}
where
\begin{equation}
\gamma= {1 \over 2}\pi \lambda v_F \omega_{2k_F}
\coth\left({\omega_{2k_F} \over 2T}\right).
\label{gamma}
\end{equation}
This model is expected to be reliable except near the transition
temperature where there is significant dispersion in the phonons.
This dispersion is now taken into account.
Near the transition temperature the
phonons can be treated {\it classically} since in most materials the
frequencies of the phonons with wavevector near $2k_F$ are much smaller than
the transition temperature (Table \ref{table2}).
Following Rice and St\"assler \cite{ric} renormalized phonon frequencies
$\Omega(q,T)$ are used in
the expression for the correlations of the random potential
\begin{equation}
\langle \Delta(z)\Delta(z')^* \rangle =
\lambda\pi T{ v_F\over a_z}
\sum_q{\omega_Q^2\over\Omega(q,T)^2}
e^{iq(z-z')}.
\label{cor}
\end{equation}
At the level of the Gaussian approximation the
phonon dispersion relation can be written in the form
\begin{equation}
\Omega(q,T)^2=\Omega(T)^2 \bigl(1 + (q- 2k_F)^2
\xi_{\parallel}(T)^2 \bigr). \label{be1}
\end{equation}
Evaluating (\ref{cor}) then gives (\ref{cor2}) where
\begin{equation}
\psi^2=\lambda \pi T
\left({\omega_Q \over \Omega(T)}\right)^2 {v_F\over 2\xi_\parallel(T)}.
\label{ce}
\end{equation}
Note that this expression togehter with (\ref{cor2})
is then quite different from (\ref{af})
used by Lee, Rice, and Anderson \cite{lee}.
In the limit $\xi_\parallel \to 0$, i.e., the phonons become
dispersionless and the sum in (\ref{cor}) becomes a delta function
and giving (\ref{corprl}) (with $\Delta_o =0$) and (\ref{gamma}).
The rms fluctuations $\delta u$ in the
positions of the atoms due to
thermal lattice motion is related to the Debye-Waller factor
and given by
\begin{equation}
(\delta u)^2=kT \sum_q {1\over M\Omega(\vec{q},T)^2} \label {ca}
\end{equation}
This is related to $\psi = (2M \omega_Q)^{1/2} g \delta u$.
Hence $\psi$ {\it is proportional to the thermal lattice motion.}
If $\psi$ is defined by (\ref{ce}) it
diverges as $T \to T_P$ because
\begin{equation}
{\rm as}\ T \to T_P,\ \Omega(T) \to 0,\ \xi_\parallel(T) \to
\infty \ {\rm with}\ \Omega(T)\xi_\parallel(T) \
{\rm finite.} \label{bf}
\end{equation}
However, in a real crystal the phonons are three-dimensional
and the thermal lattice motion is finite.
Define
\begin{equation}
\psi^2=\lambda\pi T{ v_F\over a_z}
\sum_{\vec{q}}{\omega_Q^2\over\Omega(\vec{q},T)^2} \label{cd}
\end{equation}
and write the three-dimensional dispersion relation
(for a tetragonal crystal) in the form
\begin{equation}
\Omega(\vec q,T)^2=\Omega(T)^2 \bigl(1 + (q_{\parallel}- 2k_F)^2
\xi_{\parallel}(T)^2
+ (q_{\perp}- Q_{\perp})^2 \xi_{\perp}(T)^2 \bigr) \label{be}
\end{equation}
where $\vec Q =(Q_\perp,2k_F)$ is the nesting vector associated with
the three-dimensional CDW transition (see equation (\ref{bb})).
Due to the quasi-one-dimensionality of the crystal
the dispersion perpendicular to the chains is small and
$\xi_\perp \ll \xi_\parallel$.
Let $a_x$ denote the lattice constant perpendicular to the
chains.
Performing the integral over the wave vector in (\ref{cd}) gives
\cite{alternative}
\begin{equation}
\psi^2 = \lambda \pi T
\left({\omega_Q\over\Omega(T)}\right)^2
{a_x^2 v_F \over \pi^2 \xi_\parallel(T)\xi_\perp(T)^2}
\left[\sqrt{1+(\rho_c\xi_\perp(T))^2}-1\right]
\label{cf}
\end{equation}
where $\rho_c$ is a wavevector cutoff perpendicular to the chains.
If $\rho_c=\pi/a_x$ this expression reduces to (\ref{ce}) in the
one-dimensional limit $\xi_\perp \ll a_x$.
Near the transition, $\xi_\perp(T) \to \infty$, giving
\begin{equation}
\psi(T_P)^2 = \lambda \pi T
\left({\omega_Q\over\Omega(T)}\right)^2 {a_x v_F \over \pi
\xi_\parallel(T)\xi_\perp(T) }\label{cg}
\end{equation}
From (\ref{bf}) and the fact that $\xi_\parallel(T)/\xi_\perp(T)$
is finite it follows that $\psi$ is finite as $T \to T_P$.
Note that the magnitude of this quantity is dependent on the
choice of the momentum cutoff $\rho_c$.
The above treatment is quite similar to Schulz's discussion of
fluctuations in the order parameter in the Gaussian approximation
\cite{sch}.
Although the expressions (\ref{ce}), (\ref{cf}), and
(\ref{cg}) for $\psi$ in the different regimes look very
different $\psi$ is actually weakly temperature dependent and does not
vary much in magnitude. To see this (\ref{cf}) can be written as
\begin{equation}
\psi^2 = (\psi(T_P))^2
{1 \over \rho_c\xi_\perp(T)}\left[\sqrt{1+(\rho_c\xi_\perp(T))^2}-1\right]
\label{cg2}
\end{equation}
The postfactor is a slowly varying function of $\rho_c\xi_\perp(T)$.
Since well above $T_P$, $\rho_c\xi_\perp(T) \sim 1$
(e.g., for K$_{0.3}$MoO$_3$ $\xi_\perp(300 {\rm K}) \sim 4 \AA $ \cite{gir})
the postfactor does not vary by more than a factor of two
although $\rho_c\xi_\perp(T)$ varies by several orders of
magnitude.
Johnston et al. \cite{joh} used a crude method of estimating the
pseudogap and found it to be weakly temperature dependent above $T_P$
for K$_{0.3}$MoO$_3$.
\subsection{Solution of the model}
\label{secsol}
Sadovsk\~i\~i \cite{sad2} solved the one-dimensional model (\ref{hamel})
and (\ref{cor2}) exactly. He calculated the one-electron Green's function
in terms of a continued fraction by
finding a recursion relation satisfied by the self energy.
He found \cite{sad}
that the Green's function reduced to a simple analytic form
in the limit of large correlation lengths ($\xi_\parallel \gg v_F/\psi$).
This can be seen by the following rough argument.
In the limit $\xi_\parallel \to \infty$
the moments of the random potential $\Delta(z)$ are independent of
position:
\begin{equation}
\langle \Delta(z)\rangle = 0
\ \ \ \ \ \ \langle \Delta(z)\Delta(z')^* \rangle = \psi^2.
\label{cori}
\end{equation}
This means that
the random potential has only one non-zero Fourier component, i.e.,
the one with zero wavevector.
The potential can be written $\Delta(z)=v \psi$
where $v$ is a complex random variable with a Gaussian distribution.
Averages over the random potential can then be written
\begin{equation}
\langle A[\Delta(z)]\rangle=
\int{dvdv^*\over\pi} e^{-vv*} A[v \psi].
\end{equation}
It is then a straight-forward exercise
to evaluate the averages of different electronic
Green's functions.
\section {ONE-ELECTRON GREEN'S FUNCTION NEAR $T_P$}
\label{secgreen}
The matrix Matsubara Green's function, defined at the Matsubara energies
$\epsilon_n=(2n + 1) \pi T$, for the Hamiltonian (\ref{hamel})
with (\ref{cori}) is
\begin{equation}
\hat G\left(i\epsilon_n ,k\right)=\int{dvdv^*\over\pi} e^{-vv*}
\hat G \left(i\epsilon_n,k,v\right)\label{ea}
\end{equation}
where
\begin{equation}
\hat G\left(i\epsilon_n,k,v\right)={-(i\epsilon_n
-kv_F\sigma_3-\psi(v\sigma_++v^*\sigma_-))\over\epsilon_n^2
+(kv_F)^2+vv^*\psi^2}\label{ea2}
\end{equation}
is the matrix Green's function for the Hamiltonian (\ref{hamel})
with $\Delta(z)=v\psi$.
The off-diagonal
(anomalous) terms vanish when the integral over $v$ is performed
indicating there is no long range order. The integral over
the phase of $v$ can be performed and variables
changed to $\varphi=vv^*$ and
obtain
\begin{equation}
\hat G(i\epsilon_n,k)=-(i\epsilon_n
-kv_F\sigma_3)\int_0^\infty d\varphi{e^{-\varphi}\over\epsilon_n^2
+(kv_F)^2+\varphi\psi^2}\label{eb}
\end{equation}
Sadovsk\~i\~i \cite {sad} obtained the same expression by
diagrammatic summation. For the case of a half-filled band $v$ is strictly
real and the resulting expressions are the same as those obtained by
Wonneberger and Lautenschl\"ager \cite{won}.
Expanding (\ref{eb}) in powers of $\psi$ gives
\begin{equation}
\hat G(i\epsilon_n,k)=\hat G_0(i\epsilon_n,k)\int_0^\infty d\varphi
e^{-\varphi}
\sum^\infty_{n=0}\left[{-\varphi\psi^2\over\epsilon_n^2+(kv_F)^2}
\right]^n \label{ec}
\end{equation}
where $\hat G_0=(i\epsilon_n -kv_F\sigma_3)^{-1}$
is the free-electron Green's function. Performing the
integral over $\varphi$ gives
\begin{equation}
\hat G(i\epsilon_n,k)=\hat G_0(i\epsilon_n,k)\sum^\infty_{n=0}n!
\left[{-\psi^2\over\epsilon_n^2+(kv_f)^2}\right]^n \label{ed}
\end{equation}
This is a {\it divergent} series and asymptotic expansion. However, it is
Borel summable \cite{bor}. This divergence suggests that
perturbation theory as used in References \cite {lee,cha,ric,bje,suz,sha}
may give unreliable results.
This can be seen in Figure \ref{figdos} and Section \ref{seccha}.
\subsection {Density of States}
The electronic density of states is calculated directly from the
imaginary part of the one-electron Green's function (\ref{eb}). The result is
\begin{equation}
\rho(E)=\rho_o \int_0^\infty d\varphi
e^{-\varphi}{E\over\left[E^2-\varphi\psi^2\right]^{1/2}}
\theta\left(\mid E\mid^2-\varphi\psi^2\right)
\end{equation}
\begin{equation}
=2\rho_o
\bigl|{E\over\psi}\bigl|\exp(-\left({E\over\psi}\right)^2)
{\rm erfi}\left({E\over\psi}\right)\label{rf}
\end{equation}
where $\rho_o=1/\pi v_F$ is the free-electron density of states
and erfi is the error function of imaginary argument
\cite{err}. Figure \ref{figdos}
shows the energy dependence of the density of states. It vanishes at
zero energy (the Fermi energy) and is suppressed over an energy
range of order $\psi$, i.e., there is a pseudogap.
It has the asymptotic behavior:
\begin{equation}
\rho(E) \simeq 2 \rho_o ({E\over\psi})^2 \quad {\rm for}
\ E\ll\psi \label{eg}
\end{equation}
$$\rho(E) \simeq \rho_o \quad {\rm for} \ E\gg\psi \label{eg2}$$
Figure \ref{figdos}
shows that the exact result (\ref{rf}) (solid line)
deviates significantly from the result of
second-order perturbation theory in References \cite{lee,ric} (dashed line),
\begin{equation}
\rho(E)=\rho_o {E\over\left[E^2-\psi^2\right]^{1/2}}
\theta\left(E^2-\psi^2\right) \label{eh}
\end{equation}
This latter form has been assumed in much earlier
work \cite{hor,joh,sha,joh2}.
The above expressions for the density of states
are all for infinite correlation length
($\xi_\parallel \psi/v_F \to \infty $),
i.e., very close to the three-dimensional transition temperature $T_P$.
What happens
{\it above} $T_P$ as the intrachain correlation length decreases?
This problem was considered in detail by Sadovsk\~i\~i \cite{sad2}.
(He calculated the density of states for
the random potential
(\ref{cor2}) with finite $\xi_\parallel$ exactly).
As the correlation length decreases the density of states
at the Fermi energy increases, i.e., the pseudogap fills in.
How quickly this happens depends on the dimensionless
ratio $v_F/(\psi \xi_\parallel).$
(See equation (\ref{suc}) below and Figures 5 and 6 in Reference \cite{sad2}).
Sadovsk\~i\~i showed that perturbation theory \cite {lee,cha,ric,bje,suz,sha}
only gives reliable results for $|E| < \psi$
when $\xi_\parallel < v_F/\psi$, i.e.,
well above $T_P$.
What happens
{\it below} $T_P$ as the intrachain correlation length decreases?
In Reference \cite{mck} it was shown that in the
three-dimensionally ordered Peierls state,
well below $T_P$,
there is an absolute gap with a subgap tail that
increases substantially
as the temperature becomes larger than the phonon frequency.
A smooth crossover to the pseudogap discussed here is expected.
It is an open problem to construct a single theory that
can describe the density of states over the complete
temperature range.
\subsection {Spectral Function}
The spectral function for right moving electrons
of momentum $k$ is given by
\begin{eqnarray}
A(k,E)&=&-{1\over\pi}{\rm Im}\ G_{11}(k,E+i\eta) \nonumber \\
&=&\int_0^\infty d\varphi e^{-\varphi}\left[
\delta\left(E-\sqrt{(kv_F)^2+\varphi\psi^2}\right)
+\delta\left(E+\sqrt{(kv_F)^2+\varphi\psi^2}\right)\right]
\nonumber \\
&=&{\mid E\mid\over\psi^2}\exp\left(
{(kv_F)^2-E^2\over\psi^2}\right)
\theta\left(E^2-(kv_F)^2\right) \label{ej}
\end{eqnarray}
where the momentum $k$ is relative to the Fermi momentum $k_F$.
Note that this form is very different from the Lorentzian form
associated with the quasi-particle picture and perturbation theory
\cite{rick}.
The spectral function
is asymmetrical, very broad, and has a significant high energy tail.
Figure \ref{figspec} shows how the quasi-particle weight is reduced
near the Fermi momenta, i.e., the quasi-particles are not well defined.
This was first pointed out by
Wonneberger and Lautenschl\"ager \cite{won} for the
corresponding model for a half-filled band.
This is strictly a non-perturbative effect. In perturbation theory
the quasi-particles are well defined.
This breakdown of the quasi-particle picture is similar to the
properties of a Luttinger liquid \cite{voi}.
The momentum distribution function $n(k)$ at $T=0$ for right moving
electrons is given by
\begin{equation}
n(k)\equiv \int_{-\infty}^0 dE A(k,E)
= {1\over 2} \left[ 1 - \sqrt{\pi}
\left({kv_F \over \psi} \right)
\exp \left( \left({kv_F \over \psi} \right)^2 \right)(1-
{\rm erf} \left({kv_F \over \psi} \right))
\right]
\label{ek}
\end{equation}
where ${\rm erf}$ is the error function.
The inset to Figure \ref{figspec}
shows how the momentum distribution $n(k)$
at $T=0$ is smeared over a momentum range $\delta k \sim \psi/v_F$.
The absence of a step at $k=k_F$ indicates that there
is no clearly defined Fermi surface.
However, this is {\it not} like in a Luttinger liquid,
but solely due to disorder. In fact, in an ordinary metal
with mean free path $\ell$ similar behaviour is seen;
disorder smears out $n(k)$ over a momentum range $\delta k \sim 1/\ell$.
\subsection{Electronic specific heat }
The electronic specific heat $C_e(T)$ is related to the density of
states $\rho(E)$ by
\begin{equation}
C_e(T) = - {4 \over T} \int_0^\infty dE E^2
\rho(E) {\partial f \over \partial E}
\label{spa}
\end{equation}
where $f(E)$ is the Fermi-Dirac distribution function.
In the absence of a pseudogap $C_e(T)={2 \pi^2 \over 3} \rho_0 T
\equiv C_0(T).$
If the expression (\ref{rf}) is used for the density of
states in the presence of a pseudogap then $C_e(T)/C_0(T)$
only depends on $\psi/T$ and is shown in Figure \ref{figparam}.
A similar result was recently used \cite{mck0} to explain the temperature
dependence of the electronic specific heat near a spin-density-wave
phase boundary of the organic conductor (TMTSF)$_2$ClO$_4$.
Note that when $\psi \sim T$, $C_e(T)$ can be slightly larger
than $C_0(T)$ because $E^2 {\partial f \over \partial E}$
has a maximum near $ E \sim T$ and for $ E \sim \psi$,
$\rho(E)$ is larger than $\rho_0$ (Figure \ref{figdos}).
\subsection{Pauli Spin Susceptibility}
\label{secsusc}
The Pauli spin susceptibility $\chi(T)$ is related to the density of
states $\rho(E)$ by
\begin{equation}
\chi(T) = -\mu_B^2 \int_0^\infty dE \rho(E) {\partial f \over \partial E}
\label{sua}
\end{equation}
where $f(E)$ is the Fermi-Dirac distribution function
and $\mu_B$ is a Bohr magneton \cite{ash}.
In the absence of a pseudogap $\chi(T)=\mu_B^2 \rho_0 \equiv \chi_0$
which is independent of temperature.
If the expression (\ref{rf}) is used for the density of
states in the presence of a pseudogap then $\chi(T)/\chi_0$
only depends on $\psi/T$ and is shown in Figure \ref{figparam}.
This result will be used in Section \ref{secest} to provide an
estimate of the pseudogap in K$_{0.3}$MoO$_3$.
\subsection{Chandra's scaling relation}
\label{seccha}
The effect of thermal lattice fluctuations on the temperature
dependence of $\chi(T)$ was first considered
by Lee, Rice, and Anderson \cite{lee}. They argued that as the
temperature is lowered towards $T_P$ the intrachain
correlation length increases, more of a pseudogap opens in
the density of states and $\chi(T)$ decreases.
This problem was recently reconsidered by Chandra \cite{cha}
who derived a scaling relation between the derivative
$ d \chi / d T$ and the specific heat $C_P$ in the critical region.
I now repeat the essential features of her argument.
She calculated the electronic self energy in the Born
approximation, taking into account the interchain interactions
and the finite mean free path of the electrons. She assumed
that the pseudogap is much larger than the transition
temperature ($\psi \gg T_P$; it will be shown in Section \ref{secest}
that this is poor approximation for K$_{0.3}$MoO$_3$) so that
$\chi(T) \simeq \mu_b^2\rho(0)$. Chandra also assumed that the temperature
dependence of the density of states at the Fermi energy
is determined solely by the temperature dependence
of $\xi_\parallel(T)$. Moreover, based on the Born approximation,
she found
\begin{equation}
\rho(0) \sim {1 \over \xi_\parallel(T)}.
\label{sua2}
\end{equation}
Defining $t \equiv |T-T_P|/T_P$, then gives the scaling relation
\begin{equation}
{d \chi(T) \over dT } \sim {d \over dT } {1 \over \xi_\parallel(T)}
\sim {d \over dT } t^{1/2} \sim C_P
\label{sub}
\end{equation}
where use has been made of the temperature dependence of $\xi_\parallel(T)$
and $C_P$ in the Gaussian approximation \cite{ma}.
This same scaling relation was suggested earlier
by Horn, Herman and Salamon \cite{horn}. They claimed to
have found the critical exponent for $d \chi / d T$ to be --0.5 for TTF-TCNQ.
Kwok, Gr\"uner, and Brown \cite{kwo2}
claim to have observed a scaling between
$d (T \chi) / d T$ and $C_P$ within a 30 K
region about $T_P=183$ K for K$_{0.3}$MoO$_3$.
However, Mozurkevich has argued that the
Gaussian approximation is not valid in this
temperature range \cite{moz}.
Chung {\it et al.} \cite{chu} found that $d \chi / d T$
was comparable to $C_P$ when a background contribution
was subtracted from the latter.
Brill {\it et al.} \cite{bri} found that $\chi$
was proportional to the entropy (evaluated from
integrating the specific heat) between 140 and 220 K.
(This is equivalent to
a scaling between $d \chi / d T$
and $C_P$). They show that this is
what is expected if $\chi$ and $C_P$ are derived from
a free energy functional in which the complete
magnetic field dependence is contained in the
field dependence of $T_P$.
Chandra's derivation of the scaling relationship (\ref{sub}) is
not valid. It depends on (\ref{sua2})
which is a direct result of the perturbative treatment
of the lattice fluctuations. The exact Green's function calculated
by Sadovsk\~i\~i \cite{sad2} gives different results. He found that for
$\xi_\parallel(T) \gg v_F/\psi$
\begin{equation}
{\rho(0) \over \rho_0} \simeq (0.54 \pm 0.01)
({v_F \over \psi \xi_\parallel(T)})^{1/2}
\label{suc}
\end{equation}
(see Figure 6 in Reference \cite{sad2})
rather than (\ref{sua2}).
This will give
\begin{equation}
{d \chi(T) \over dT } \sim t^{-3/4}
\label{sud}
\end{equation}
and so the scaling relation (\ref{sub}) does not hold.
It should be stressed that this result assumes $\psi \gg T_P$,
a condition that is poorly satisfied in most materials
(Section \ref{secest}).
\section{PROPERTIES OF THE GINZBURG-LANDAU COEFFICIENTS}
\label{secprop}
In Appendix \ref{seccoeff} the coefficients $a$, $b$, and $c$ in the
Ginzburg-Landau free energy (\ref{aa1}) describing the
Peierls transition are evaluated in the presence of the
random potential (\ref{cori}) which is used here to model the
thermal lattice motion. The calculation is based on a linked
cluster expansion similar to that used to derive
the Ginzburg-Landau functional for superconductors \cite{has}.
The results are:
\begin{equation}
a(T)= {1 \over \pi v_F} \left[
\ln\left({T\over T_{RL}}\right)+\pi
T\sum_{\epsilon_n}\left({1\over\left|\epsilon_n\right|}-\int_0^\infty
d\varphi
e^{-\varphi}{\epsilon_n^2\over\left(\epsilon_n^2+
\varphi \psi^2\right)^{3/2}}\right) \right]
\label{gla}
\end{equation}
\begin{equation}
b(T)={ T \over 4 v_F}\sum_{\epsilon_n}
\int_0^\infty d\varphi e^{-\varphi}
\left( {\epsilon_n^2 \over
\left(\epsilon_n^2+ \varphi\psi^2 \right)^{5/2}}
-{5 \varphi(\psi \ \epsilon_n)^2 \over
\left(\epsilon_n^2+ \varphi \psi^2 \right)^{7/2}}
\right)
\label{glb}
\end{equation}
\begin{equation}
c(T)={v_F T\over 4}
\sum_{\epsilon_n}\epsilon_n^2\int_0^\infty{d\varphi
e^{-\varphi} \over (\epsilon_n^2+\varphi\psi^2)^{5/2}}\label{ff}
\label{glc}
\end{equation}
The integrals over $\varphi$ in the above
expressions can be written in terms of error functions and incomplete
gamma functions \cite{err}. However, for both numerical and
analytical calculations it is actually more convenient to
use the expressions above.
As $\psi \to 0$ the above expressions
reduce to the rigid-lattice values (\ref{mfa}--\ref{mfc}).
{\it Single-chain mean-field transition temperature.}
$T_0$ is determined by the temperature
at which the second-order Ginzburg-Landau coefficient (\ref{gla}) vanishes:
\begin{equation}
a(T_0)=0. \label{fk2}
\end{equation}
This defines relations between $T_0/T_{RL}$ and
$\psi/T_0$, shown in Figure \ref{figcoeff}
(The inset shows $T_0/T_{RL}$ versus $\psi/T_{RL}$).
The pseudogap
suppresses the transition temperature. At a crude level, this is
because in the presence of a pseudogap
opening a gap due to a Peierls distortion causes a smaller
decrease in the electronic energy than in the absence of a pseudogap.
In most materials $T_P < 0.4 T_{RL}$ (Table \ref{table2})
and so the inset of Figure \ref{figcoeff} implies $\psi \sim T_{RL}$
which is comparable to the zero-temperature gap.
Rice and Str\"assler \cite{ric} found from second-order perturbation theory
that for $T_0 \ll T_{RL},$ $\psi \simeq 1.05 T_{RL}.$
Thus, the single-chain mean-field transition temperature
can be quite different from
$T_{RL}$, defined by (\ref{trl}), and often referred to
as the mean-field transition temperature,
and so no experimental signatures are expected at $T=T_{RL}$.
{\it Fourth-order coefficient.}
The ratio of the fourth-order coefficient $b$ to its rigid-lattice value
as a function of the ratio of the pseudogap $\psi$ to the
temperature is shown in Figure \ref{figcoeff}.
Note that $b$ is negative for $\psi/T > 2.7$.
This will change the nature of the phase transition.
One must then include the sixth-order term in the free energy.
If it is positive (I have calculated it and found it to be positive
for this parameter range)
then the transition will be {\it first order.}
A complete discussion of such a situation is given by
Toledano and Toledano \cite{tol}.
Imry and Scalapino have discussed the effect of
one-dimensional fluctuations for this situation \cite{imr}.
At the mean-field level there is a co-existence of phases
for the temperature range defined by
\begin{equation}
0 < a(T) < { b(T)^2 \over 3 d(T)}
\label{fm0}
\end{equation}
where $d(T)$ is the sixth-order coefficient.
Hysteresis will be observed
in this temperature range.
I recently suggested that the first-order nature of the destruction
by high magnetic fields of spin-density-wave states
in organic conductors is due to similar effects \cite{mck0}.
If at low temperatures the electron phonon coupling
$\lambda$ is varied then $\psi/T_{RL} \sim \lambda e^{1/\lambda}$.
According to the inset of Figure \ref{figcoeff} there will
be a critical coupling below which the CDW phase will
be destroyed. This transition will be first order.
It is interesting that Altshuler, Ioffe, and Millis \cite{alt}
recently obtained a similar result for a two-dimensional
Fermi liquid (with a quasi-one-dimensional
Fermi surface) using a very different approach.
However, it should be pointed out that when $b$ is small
corrections due to other effects such as a finite correlation
length and interchain coupling
will be important and could make $b$ positive.
It is unclear whether this unexpected behaviour is only a result
of the simplicity of the model or actually is relevant to
real materials. The
three-dimensional transition occurs when the parameter $\kappa$,
defined by (\ref{aat1}), becomes sufficiently small \cite{mck2}.
Generally this is assumed to be due to the temperature becoming sufficiently
low. However, I speculate that the transition
could alternatively be driven by $b$ becoming
sufficiently small.
The fact that $\psi \sim (2-3) T_P$ in K$_{0.3}$MoO$_3$
(Section \ref{secest}) is consistent with $b$ being small.
{\it The coefficient of the longitudinal gradient term}
is given by (\ref{glc}).
It can be shown that $c(T)/c^{RL}(T)$ is a universal
function of $\psi/T$ (see Figure \ref{figcoeff})
and that the pseudogap reduces the
value of $c$.
{\it Interchain coupling.}
Consider a crystal with tetragonal unit cell of dimensions
$a_x \times a_x \times a_z$,
where the z-axis is parallel to the chains. For a tight-binding model
the electronic band structure is given by the dispersion relation
\begin{equation}
E(k)=-2t_\perp(\cos(k_xa_x)+\cos(k_ya_x))-2t_\parallel \cos(k_z a_z).
\label{ba}
\end{equation}
Assume the band-structure is highly anisotropic, i.e., $t_\parallel\gg
t_\perp$.
The Fermi velocity $v_F$ is defined by $v_F=2t_\parallel a_z \sin(k_F a_z)$.
Horovitz, Gutfreund, and Weger \cite {hor} have shown that
imperfect nesting of the Fermi surface (i.e., $E(k) \simeq - E(k+Q))$
occurs for the nesting vector
\begin{equation}
\vec{Q}=(\pi/a_x,\pi/a_x,2k_F). \label {bb}
\end{equation}
To calculate the interchain coupling $J$ in the
Ginzburg-Landau functional (\ref{ad1}) it is assumed that
the one-dimensional Green's function (\ref{ea2})
can simply be replaced
with the corresponding one with the anisotropic band structure, given by
equation (\ref{ba}).
The calculation
is then essentially identical to the rigid-lattice calculation of Horovitz,
Gutfreund, and Weger \cite{hor} and so only the result is given
(compare (\ref{mfc2})):
\begin{equation}
J = \left( { 4 t_\perp \over v_F}\right)^2 c(T).
\label{fk}
\end{equation}
Since the pseudogap reduces the value of the longitudinal
coefficient $c$ it will also reduce the interchain coupling.
\section{MEAN-FIELD THEORY OF A SINGLE CHAIN}
The single chain Ginzburg-Landau functional
with the coefficients discussed in the previous section
is now considered.
In particular it is shown that the one-dimensional fluctuations
can be much smaller than for the functional with the
rigid-lattice coefficients.
The first step is to consider the temperature dependence of
the second-order coefficient $a(T)$ near $T_0$,
the mean-field transition temperature.
This is difficult because to be realistic the
temperature dependence of the parameter $\psi$
must be included. This is done at a crude level
using the simple model based on the discussion of
thermal lattice motion in Section \ref{sectherm}.
This is then used to evaluate $a^\prime$, defined
by (\ref{aa10}), and needed to evaluate physical
quantities associated with the transition:
the specific heat jump, the coherence length,
and width of the critical region.
The jump in the specific heat at $T_0$ is
\begin{equation}
\Delta C_{1D}=
{(a^\prime)^2 \over 2 \ b \ T_0}.
\label{ac1}
\end{equation}
An important length scale is the coherence length $\xi_0$,
defined by
\begin{equation}
\xi_0=\left({c \over a^\prime}\right)^{1/2}
\label{acd1}
\end{equation}
The one-dimensional
Ginzburg criterion \cite{gin} provides an estimate of the
temperature range, $\Delta T_{1D}$, over which critical fluctuations
are important.
\begin{equation}
\Delta t_{1D} \equiv {\Delta T_{1D} \over T_0}
= \left({b \ T_0 \over a^{\prime 3/2} c^{1/2}}
\right)^{2/3}
= {1 \over (2 \xi_0 \Delta C_{1D})^{2/3}}
\label{acc1}
\end{equation}
\subsection {Self-consistent determination of the pseudogap}
At the level of the Gaussian approximation the
phonon dispersion is related to the Ginzburg-Landau
coefficients by
\begin{equation}
\Omega(q,T)^2=\lambda \omega_Q^2 \bigl(a(T) +
c(T) (q- 2k_F)^2 + Ja_x^2 (q_\perp - Q_\perp)^2 \bigr). \label{bz1}
\end{equation}
Hence the phonon dispersion depends on the pseudogap
$\psi$.
However, it was shown in Section \ref{sectherm} that $\psi$ depends on the
dispersion. Hence, $\psi$ must be determined self-consistently.
Equation (\ref{cg}) gives the dependence of the pseudogap at $T_0$ on the
phonon dispersion. Equation (\ref{ff}) gives the dependence of the
coefficient $c(T)$ on the pseudogap.
These can be combined with (\ref{fk}) to give
\begin{equation}
1= t_\perp \psi^2 \
\sum_{\epsilon_n}
\epsilon_n^2\int_0^\infty {d\varphi e^{-\varphi}\over
(\epsilon_n^2+\varphi\psi^2)^{5/2}}.\label{fm}
\end{equation}
It follows that $\psi/T$ is a universal function of
$t_\perp /T$.
{\it Dependence of $T_0$ on the interchain interactions.}
The self-consistent equation for the pseudogap (\ref{fm})
can be solved simultaneously with the equations for $T_0$,
and (\ref{fk})
to give the single-chain mean-field
transition temperature as function of the interchain interactions.
The transition temperature is then a monotonic
increasing function of the interchain hopping.
A similar procedure was followed by Rice and Str\"assler
\cite{ric}.
The transition temperature tends to zero as the interchain
coupling tends to zero, consistent with the fact that there
are no finite temperature phase transitions in a strictly one-dimensional
system \cite{lan}.
\subsection {Evaluation of $a'$}
It is now assumed that the temperature dependence of
the pseudogap $\psi$ is
given implicitly by equation
(\ref{fm}).
Implicit differentiation then gives
\begin{equation}
{d \over dT} \left( {\psi \over T} \right)=
{\psi \over 2 T^2}
{X(T) \over Y(T)}
\label{yz}
\end{equation}
where
\begin{equation}
X(T)=\sum_{\epsilon_n}\epsilon_n^2\int_0^\infty{d\varphi
e^{-\varphi} \over (\epsilon_n^2+\varphi\psi^2)^{5/2}}
\end{equation}
\begin{equation}
Y(T)=\sum_{\epsilon_n}\epsilon_n^2\int_0^\infty{d\varphi
e^{-\varphi} \varphi \over (\epsilon_n^2+\varphi\psi^2)^{5/2}}
\end{equation}
Note that since the right-hand side of (\ref{yz})
is positive
$\psi/T$ is always an increasing function of temperature.
A lengthy calculation gives
\begin{equation}
a'={1 \over \pi v_F} \left( 1 + {3 \over 2} \psi^2 \pi T
\sum_{\epsilon_n}\epsilon_n^2\int_0^\infty{d\varphi
e^{-\varphi} \over (\epsilon_n^2+\varphi\psi^2)^{5/2}}
\right)
\end{equation}
This is large than the rigid-lattice value
$a'_{RL} \equiv 1/\pi v_F$. This enhancement
will enhance the specific heat jump (\ref{ac1}) and reduce the
coherence length (\ref{acd1}).
\subsection {Specific heat jump}
The specific heat jump $\Delta C$ at the transition temperature
is calculated from equation (\ref{ac1}). It is shown in Figure
\ref{figjump}.
Note that the jump is much larger than the
rigid-lattice value of $1.43\gamma T_P$.
The trend shown in Figure \ref{figjump} can be explained by a
rough argument correlating the sizes of $\Delta C/ \gamma T_P$
and $\Delta(0)/k_B T_P$. Simply put, if $\Delta(0)/k_B T_P$ is large
then $\Delta(T)^2$ will have a large slope at $T_P$.
It has previously been noted
experimentally \cite{sat} that the order parameter has a
BCS temperature dependence with $\Delta(0)$
and $T_P$ treated as independent parameters.
Some theoretical justification was recently provided
for such a temperature dependence well away from $T_P$ \cite{mck}.
Close to $T_P$ the BCS form gives
\begin{equation}
\Delta(T) \simeq 1.74 \Delta(0)\left(1-{T \over T_P}\right)^{1/2}.
\end{equation}
Within a BCS type of framework
the specific heat discontinuity is given by \cite{tin}
\begin{equation}
\Delta C \sim -\rho_o {d \Delta^2 \over dT} \Big|_{T_P}
= 3.03 \rho_o {\Delta(0) ^2 \over T_P}.
\end{equation}
Using $\Delta(0)=1.76k_B T_{RL}$ and $\gamma= 2 \pi^2 \rho_o/3$
gives
\begin{equation}
{\Delta C \over 1.43 \gamma T_P} \sim \left({ T_{RL} \over T_P}\right)^2
\end{equation}
This simple argument gives the correct trend that as the
fluctuations increase the enhancement of the specific heat jump
increases.
\subsection {Width of the one-dimensional critical region}
The width of the one-dimensional critical region
is calculated from equation (\ref{acc1})
with the Ginzburg-Landau coefficients in the
presence of the pseudogap. It is shown in Figure
\ref{figjump}, normalized to the rigid-lattice value
$\Delta t_{1D}=0.8$.
The large reduction is very important because it
means that even for weak interchain coupling,
it may be possible for condition (\ref{ad10}) to
be satisfied and for a mean-field treatment of
a single chain functional, such as that used in
this section, to be justified.
\section{Estimate of the pseudogap in K$_{0.3}{\rm MoO}_3$}
\label{secest}
Optical conductivity, magnetic susceptibility and
photoemission experiments all suggest that near $T_P = 183$ K
there is a pseudogap in the density of states.
{\it Optical conductivity}. Sadovsk\~i\~i has calculated the
optical conductivity $\sigma(\omega)$ for the model introduced
in Section \ref{secham}
\cite{sad}. For small frequencies $\sigma(\omega)$
is linear in $\omega$ and has a peak at about $\omega \simeq 3 \psi$.
The data in References \cite{deg,dre} then implies
$\psi \sim $ 40 meV
and $\psi/T_P \sim 2.5$.
On a less rigorous level $\psi$ can be estimated based on the
analysis contained in the inset of Figure \ref{figcoeff}. If
the single-chain mean-field transition temperature $T_0 < 0.4 T_{RL}$
then $\psi \sim T_{RL}$. Using the BCS relation (\ref{trl})
and the estimate $\Delta (0) \simeq 80 $ meV for the zero-temperature
gap from the optical conductivity \cite{deg} gives
$\psi \sim $ 45 meV and $\psi/T_P \sim 3$.
{\it Magnetic susceptibility.}
The data of References \cite{bri,joh} gives $\chi(T_P)/\chi(300 K)
\simeq 0.5 $.
Assuming that $\chi(300 {\rm K}) \simeq \chi_0$
and using Figure \ref{figparam} gives $\psi /T_P \sim 2.4$.
Note that all of the above three estimates for $\psi/T_P$
are consistent with one another and
are all in the regime where the fourth-order coefficent $b$
is small (Figure \ref{figcoeff}).
{\it Photoemission.}
Recent high resolution photoemssion measurements
\cite{dar,dar2,hwu}
on K$_{0.3}$MoO$_3$ and (TaSe$_4$)$_2$I
have several puzzling features:
(1) There is a suppression of spectral weight over a large energy range
(of the order of 200 meV for K$_{0.3}$MoO$_3$)
near the Fermi energy.
(2) The spectrum is very weakly temperature dependent. The suppression
occurs even for $T \sim 2 T_P$.
(3) At $T_P$ the spectrum does not just shift near $E_F$, due to the
opening of the Peierls gap, but also at energies of order 0.5 eV
from $E_F.$
These features {\it cannot} be explained using the model
presented in this paper.
The photoemission data suggests that the pseudogap is
about $\psi \sim $ 130 meV. Clearly this estimate is inconsistent
with the estimates ($\psi \sim $ 40-50 meV)
given above from the optical conductivity
and magnetic susceptibility.
Furthermore, in the model presented here
the pseudogap occurs only when $\xi_\parallel(T) \gg v_F/\psi$,
i.e., fairly close to $T_P$.
The temperature dependence of the Pauli spin susceptibility
and the optical conductivity \cite{deg,deg2} suggest that
the pseudogap disappears for $T ~> 2 T_P$ (in contrast to
(2) above). Dardel et al. \cite{dar} speculate that the
anomalous behaviour that they observe may arise
because the photoemission intensity $I(E)$ might be related
to the density of states $\rho(E)$ by $I(E)=Z \rho(E)$
and the quasi-particle weight $Z$ vanishes
due to Luttinger liquid effects.
This suggestion has been examined critically by Voit \cite{voi}
who concludes
that the photoemission data is only quantitatively
consistent with a Luttinger
liquid picture if very strong long-range interactions are involved.
Kopietz, Meden, and Sch\"onhammer \cite{kop}
have recently considered such models.
\section{CONCLUSIONS}
In this paper a simple model has been used to illustrate
some of the difficulties involved in constructing from
microscopic theory a Ginzburg-Landau theory of the CDW transition.
The main results are:
(1) The large thermal lattice motion near the transition temperature
produces a pseudogap in the density of states.
(2) Perturbation theory diverges and gives unreliable results.
This is illustrated by showing that a predicted \cite{cha} scaling relation
between the specific heat and the temperature derivative
of the susceptibility does not hold.
(3) The pseudogap significantly alters the coefficients in the
Ginzburg-Landau free energy.
The result is that one-dimensional order parameter
fluctuations are less important,
making a mean-field treatment of the
single-chain Ginzburg-Landau functional more reasonable.
This work raises a number of questions and opportunities for
future work.
(a) The most important problem is that there is still no
microscopic theory that can make reliable quantitative
predictions about how dimensionless ratios such as
$\Delta (0)/k_B T_P$, $\Delta C/ \gamma T_P$, and $\xi_{0z}T_P/v_F$
depend on parameters such as $v_F$, the electron phonon coupling
$\lambda$, $T_P$ and the interchain coupling.
(b) Is the change of the sign of the fourth-order
coefficient $b$ of the single-chain Ginzburg-Landau functional
for $\psi > 2.7 T_P$
an important physical effect or merely a result of
the simplicity of the model considered here?
(c) Calculation of the contribution of the sliding CDW
to the optical conductivity in the presence of the short-range
order associated with the pseudogap \cite{dre,dre2}.
\acknowledgments
I have benefitted from numerous discussions with J. W. Wilkins.
This work was stimulated by discussions with J. W. Brill.
I am grateful to him for showing me his group's data
prior to publication.
I thank K. Bedell and K. Kim for helpful discussions.
Some of this work was performed at The Ohio State University
and supported by the U.S. Department of Energy,
Basic Energy Sciences, Division of Materials Science
and the OSU Center for Materials Research.
Work at UNSW was supported by the Australian Research Council.
\twocolumn
\narrowtext
\begin {references}
\bibitem[*]{email}electronic address: <EMAIL>
\bibitem{gru} G. Gr\"uner, {\it Density Waves in Solids},
(Addison-Wesley, Redwood City, 1994).
\bibitem{con} For a review, see {\it Highly Conducting Quasi-One-Dimensional
Organic Crystals},
edited by E. Conwell (Academic, San Diego, 1988).
\bibitem{gor} For a review, see {\it Charge Density Waves in Solids}, edited by
L. P. Gorkov and G. Gr\"uner (North-Holland, Amsterdam, 1989).
\bibitem{car} K. Carneiro, in {\it Electronic Properties of Inorganic
Quasi-One-Dimensional Compounds, Part 1}, edited by P. Monceau (Reidel,
Dordrecht, 1985), p.1.
\bibitem{bri} J. W. Brill, M. Chung, Y.-K. Kuo, X. Zhan,
E. Figueroa, and G. Mozurkewich, Phys. Rev. Lett.
{\bf 74}, 1182 (1995);
and references therein.
\bibitem{gir} S. Girault, A. H. Moudden, and J. P. Pouget,
Phys. Rev. {\bf 39}, 4430 (1989).
\bibitem{gin} V. L.Ginzburg, Fiz. Tverd. Tela
{\bf 2}, 2031 (1960) [Sov. Phys. Solid State {\bf 2}, 1824
(1960)].
The width of the critical region, $\Delta T$, is defined by the
temperature at which the fluctuation contribution
to the specific heat below the transition temperature,
calculated in the Gaussian
approximation, equals the mean-field specific heat jump $\Delta C$.
It should be stressed that this gives only a very rough
estimate of the importance of fluctuations and that
there are several alternative definitions of the width
of the critical region.
Consequently, care should be taken when comparing estimates
from different references. This is particulary true
since different definitions can differ by
numerical factors as large as $32 \pi^2$!
\bibitem{schr} J. R. Schrieffer, {\it Theory of Superconductivity},
(Addison-Wesley, Redwood City, 1983) Revised edition, p. 248 ff.
\bibitem{carb} See e.g., J. P. Carbotte, Rev. Mod. Phys.
{\bf 62}, 1027 (1990).
\bibitem{mck2}R. H. McKenzie, Phys. Rev. B
{\bf 51}, 6249 (1995).
\bibitem{lan} L. D. Landau and E. M. Lifshitz, {\it Statistical
Physics}, 2nd. ed., (Pergamon, Oxford, 1969), p. 478.
\bibitem{sca} D. J. Scalapino, M. Sears, and R. A. Ferrell, Phys. Rev. B
{\bf 6}, 3409 (1972).
\bibitem{fro} H. Fr\"ohlich, Proc. R. Soc. London A {\bf 223}, 296
(1954).
\bibitem{ric0}M. J. Rice and S. Str\"assler,
Solid State Commun. {\bf 13}, 125 (1973).
\bibitem{all} D. Allender, J. W. Bray, and J. Bardeen, Phys. Rev. B
{\bf 9}, 119 (1974).
\bibitem{sch} H. J. Schulz, in {\it Low-Dimensional Conductors and
Superconductors}, edited by D. J\'erome and L.G. Caron (Plenum, New York,
1986), p. 95.
\bibitem {hor} B. Horovitz, H. Gutfreund, and M. Weger,
Phys. Rev. B {\bf 12}, 3174 (1975).
\bibitem{ma} S. K. Ma, {\it Modern Theory of Critical Phenomena},
(Benjamin/Cummings, Reading, 1976), p.94.
\bibitem{sco} J. C. Scott, S. Etemad, and E. M. Engler, Phys. Rev. B
{\bf 17}, 2269 (1978) [TSeF-TCNQ, TTF-TCNQ].
\bibitem{joh} D. C. Johnston, Phys. Rev. Lett. {\bf 52}, 2049
(1984) [K$_{0.3}$Mo0$_3$].
\bibitem{joh3} D. C. Johnston, M. Maki, and G. Gr\"uner,
Solid State Commun. {\bf 53}, 5 (1985) [(TaSe$_4$)$_2$I].
\bibitem{deg} L. Degiorgi, G. Gr\"uner, K. Kim, R.H. McKenzie and P.
Wachter, Phys. Rev. B {\bf 49}, 14754 (1994) [K$_{0.3}$Mo0$_3$].
\bibitem{deg2} L. Degiorgi, St. Thieme, B. Alavi,
G. Gr\"uner, R.H. McKenzie, K. Kim, and F. Levy,
Phys. Rev. B, to appear (1995)
[K$_{0.3}$Mo0$_3$, (TaSe$_4$)$_2$I].
\bibitem{dre} B. P. Gorshunov, A. A. Volkov, G. V. Kozlov,
L. Degiorgi, A. Blank, T. Csiba,
M.Dressel, Y. Kim, A. Schwartz, and G. Gr\"uner,
Phys. Rev. Lett. {\bf 73}, 308 (1994) [K$_{0.3}$Mo0$_3$].
\bibitem{dre2}
A. Schwartz, M.Dressel, B. Alavi, S. Dubois, G. Gr\"uner,
B. P. Gorshunov, A. A. Volkov, G. V. Kozlov,
S. Thieme, L. Degiorgi, and F. L\'evy,
Phys. Rev. B, to appear (1995) [K$_{0.3}$Mo0$_3$].
\bibitem{bru} P. Br\"uesch, S. Str\"assler, and H. R. Zeller, Phys. Rev. B
{\bf 12}, 219 (1975) [K$_2$Pt(CN)$_4$Br$_{0.3}$].
\bibitem{ber} D. Berner, G. Scheiber, A. Gaymann, H. P. Geserich,
P. Monceau, and F. L\'evy, J. Phys. France IV,
{\bf 3}, 255 (1993) [(TaSe$_4$)$_2$I].
\bibitem{dar} B. Dardel, D. Malterre, M. Grioni, P. Weibel, and Y. Baer,
Phys. Rev. Lett. {\bf 61}, 3144 (1991)
[K$_{0.3}$Mo0$_3$, (TaSe$_4$)$_2$I].
\bibitem {dar2} B. Dardel, D. Malterre, M. Grioni, P. Weibel, Y. Baer,
C. Schlenker, and Y. P\'etroff, Europhys. Lett. {\bf 19}, 525
(1992) [K$_{0.3}$Mo0$_3$].
\bibitem{hwu} Y. Hwu, P. Alm\'eras, M. Marsi, H. Berger,
F. L\'evy, M. Grioni, D. Malterre, and G. Margaritondo,
Phys. Rev. B {\bf 46}, 13624 (1992).
\bibitem{mck}R. H. McKenzie and J. W. Wilkins,
Phys. Rev. Lett. {\bf 69}, 1085 (1992).
\bibitem{kim}K. Kim, R. H. McKenzie, and J. W. Wilkins,
Phys. Rev. Lett. {\bf 71}, 4015 (1993).
\bibitem{lon} F. H. Long, S. P. Love, B. I. Swanson, and R. H. McKenzie,
Phys. Rev. Lett. {\bf 71}, 762 (1993).
\bibitem{lee}P. A. Lee, T. M. Rice, and P. W. Anderson,
Phys. Rev. Lett. {\bf 31}, 462 (1973).
\bibitem{die} W. Dieterich, Adv. Phys. {\bf 25}, 615 (1976).
\bibitem{sca2} D. J. Scalapino, Y. Imry, and P. Pincus, Phys. Rev. B
{\bf 11}, 2042 (1975).
\bibitem{sad} M. V. Sadovsk\~i\~i, Zh. Eksp. Teor. Fiz. {\bf 66}, 1720
(1974) [Sov. Phys. JETP {\bf 39}, 845 (1974)]; Fiz. Tverd. Tela
{\bf 16}, 2504 (1974) [Sov. Phys. Solid State {\bf 16}, 1632
(1975)].
\bibitem{mck0} R. H. McKenzie,
Phys. Rev. Lett. {\bf 74}, 5140 (1995).
\bibitem{voi} J. Voit, J. Phys. Condens. Matter {\bf 5}, 8305 (1993)
and references therein.
\bibitem{cha}P. Chandra, J. Phys. Condens. Matter {\bf 1}, 10067 (1989).
\bibitem{ric}M. J. Rice and S. Str\"assler,
Solid State Commun. {\bf 13}, 1389 (1973).
\bibitem{bje} A. Bjeli\~s and S. Bari\~si\'c, J. Physique Lett.
{\bf 36}, L169 (1975).
\bibitem{suz} Y. Suzumura and Y. Kurihara, Prog. Theor. Phys.
{\bf 53}, 1233 (1975).
\bibitem{sha} L. J. Sham, in {\it
Highly conducting one-dimensional
solids}, edited by J. T. Devreese, R. P. Evrard, and V. E. Van Doren
(Plenum, New York, 1979), p. 277.
\bibitem{bra}S. A. Brazovskii and I. E. Dzyaloshinskii,
Zh. Eksp. Teor. Fiz. {\bf 71}, 2338 (1976)
[Sov. Phys. JETP {\bf 44}, 1233 (1976)].
\bibitem{mck1}R. H. McKenzie and J. W. Wilkins,
Synth. Met. {\bf 55-57}, 4296 (1993).
\bibitem{alternative}
There are different ways of handling the cutoff.
In evaluating this integral I have followed Rice
and Str\"assler \cite{ric} and Schulz \cite{sch}
and performed the integral over $q_\parallel$
without any cutoff while a cutoff is used for
$q_\perp$. A slightly different result is
obtained if a cutoff is included also for $q_\parallel$.
\bibitem{sad2} M. V. Sadovsk\~i\~i, Zh. Eksp. Teor. Fiz. {\bf 77}, 2070
(1979) [Sov. Phys. JETP {\bf 50}, 989 (1979)].
\bibitem{won} W. Wonneberger and R. Lautenschl\"ager,
J. Phys. C. {\bf 9}, 2865 (1976).
\bibitem{bor} For a definition and discussion of Borel summation see
J. W. Negele and H. Orland, {\it Quantum Many-Particle Systems},
(Addison Wesley, Redwood City, 1988), p. 373.
\bibitem{err} M. Abramowitz and I. A. Stegun,
{\it Handbook of Mathematical Functions},
(Dover, New York, 1972). This integral is also
known as Dawson's integral.
\bibitem{joh2} D. C. Johnston, Solid State Commun. {\bf 56}, 439
(1985); D. C. Johnston, J. P. Stokes, and R. A. Klemm,
J. Mag. Mag. Mat. {\bf 54-57}, 1317 (1986).
\bibitem{rick}G. Rickayzen, {\it Green's Functions and Condensed
Matter}, (Academic, London, 1984), p.37.
\bibitem{ash}N. W. Ashcroft and N. D. Mermin,
{\it Solid State Physics} (Saunders, Philadelphia, 1976), p. 663.
\bibitem{horn}P. M. Horn, R. Herman, and M. B. Salamon,
Phys. Rev. B {\bf 16}, 5012 (1977).
\bibitem{kwo2} R. S. Kwok, G. Gr\"uner, and S. E. Brown, Phys. Rev. Lett.
{\bf 65}, 365 (1990)
[K$_{0.3}$Mo0$_3$].
\bibitem{moz} G. Mozurkewich,
Phys. Rev. Lett. {\bf 66}, 1645 (1991);
R. S. Kwok, G. Gr\"uner, and S. E. Brown,
ibid. {\bf 66}, 1646 (1991).
\bibitem{chu} M. Chung, Y.-K. Kuo, G. Mozurkewich, E. Figueroa,
Z. Teweldemedhin, D. A. Dicarlo, M. Greenblatt, and
J. W. Brill, J. Phys. France IV {\bf 3}, 247 (1993)
[K$_{0.3}$Mo0$_3$].
\bibitem{has} R. F. Hassing and J. W. Wilkins,
Phys. Rev. B {\bf 7}, 1890 (1973).
\bibitem{tol} J. C. Toledano and P. Toledano, {\it The Landau theory
of phase transitions: application to structural, incommensurate, magnetic,
and liquid crystal systems}, (World Scientific, Singapore, 1987), p. 167.
\bibitem{imr}
Y. Imry and D. J. Scalapino, Phys. Rev. B
{\bf 9}, 1672 (1974).
\bibitem{alt}
B. L. Altshuler, L. B. Ioffe, and A. J. Millis,
preprint, cond-mat/9504024
\bibitem{sat} See e.g., M. Sato, M. Fujishita, S. Sato, and S. Hoshino,
J. Phys. C {\bf 18}, 2603 (1985); R. M. Fleming, L. F. Schneemeyer,
and D. E. Moncton, Phys. Rev. B {\bf 31}, 899 (1985).
\bibitem{tin} M. Tinkham, {\it Introduction to Superconductivity},
(Krieger, Malabar, 1985), p.36.
\bibitem{kop} P. Kopietz, V. Meden, K. Sch\"onhammer,
Phys. Rev. Lett. {\bf 74}, 2997 (1995).
\bibitem{wha} M. -H. Whangbo and L. F. Schneemeyer, Inorg. Chem. {\bf
25}, 2424 (1986).
\bibitem{agd} A. Abrikosov, L. P. Gorkov, and I. E. Dzyaloshinskii,
{\it Methods of Quantum Field Theory in Statistical Physics},
(Dover, New York, 1975), p. 130.
\end{references}
\narrowtext
\twocolumn
\centerline{\epsfxsize=7.0cm \epsfbox{fig1.ps}}
\begin{figure}
\caption{Pseudo-gap in the density of states near the
three-dimensional transition
temperature $T_P$. Perturbative treatments (dotted line,
compare Ref. \protect\cite{lee,ric}) give an absolute gap
$\psi$ at the transition temperature
whereas the exact treatment (solid line) gives
only a pseudogap. The energy $E$ is relative to the
Fermi energy and the density of states is normalized
to the free-electron value $\rho_o$.
The density of states is symmetrical about $E=0$.
This result is only valid sufficiently close to
$T_P$ that the longitudinal CDW correlation length
$\xi_\parallel \gg v_F/\psi$. As the temperature
increases above $T_P$, $\xi_\parallel $ decreases and
the density of states at the Fermi energy increases, i.e.,
the pseudogap gradually fills in (see Figures 5 and 6
in Reference \protect\cite{sad2}).
\label{figdos}}
\end{figure}
\centerline{\epsfxsize=7.0cm \epsfbox{fig2.ps}}
\begin{figure}
\caption{Breakdown of the quasi-particle picture. The electronic
spectral function is shown for several different momenta $k$,
relative to the Fermi momentum $k_F$.
As the momentum approaches $k_F$
the spectral function broadens
significantly, similar to the behaviour of a Luttinger liquid.
Inset: Momentum dependence of the occupation function $n(k)$.
The dashed line is the result in the absence of a pseudogap,
i.e., a non-interacting Fermi gas.
\label{figspec}}
\end{figure}
\centerline{\epsfxsize=7.5cm \epsfbox{fig3.ps}}
\begin{figure}
\caption{
Modification of the electronic specific heat $C_e(T)$ and
the Pauli spin susceptibility $\chi(T)$
by the pseudogap.
Both are normalized to their values in the
absence of the pseudogap.
\label{figparam}}
\end{figure}
\centerline{\epsfxsize=7.5cm \epsfbox{fig4.ps}}
\begin{figure}
\caption{The pseudogap due to thermal lattice motion has
a significant effect on the coefficients in the
Ginzburg-Landau free energy (\protect\ref{aa1}) for a single chain.
The ratio of the single-chain
mean-field transition temperature $T_0$
and the coefficients $b$ and $c$ to
their rigid lattice values (given by (\protect\ref{mfa} -
\protect\ref{mfc})) are shown as a function of
the ratio of the pseudogap $\psi$ to the temperature.
For $\psi > 2.7 T$ the coefficient $b$ becomes
negative and the transition will be first order
(Section \protect\ref{secprop}).
Inset: Relationship between $T_0/T_{RL}$ and $\psi/T_{RL}$.
\label{figcoeff}}
\end{figure}
\centerline{\epsfxsize=7.5cm \epsfbox{fig5.ps}}
\begin{figure}
\caption{
Dependence on the pseudogap of physical quantities
associated with mean-field theory of a single chain.
The plot shows the coherence length $\xi_0$,
the width of the one-dimensional critical region $\Delta t_{1D}$,
and the inverse of the specific heat jump $\Delta C$.
All quantities are normalized to their rigid-lattice values.
For $\psi > 2.7 T$ the coefficient $b$ becomes
negative and the transition will be first order
(Section \protect\ref{secprop}).
The large reduction of $\Delta t_{1D}$ below the rigid lattice
value of 0.8 means that a mean-field treatment of the single
chain Ginzburg-Landau functional may be justified.
\label{figjump}}
\end{figure}
\begin{table}
\caption{
Comparison of experimental values for
K$_{0.3}$MoO$_3$ of various dimensionless ratios
with the predictions of two simple microscopic models.
The three-dimensional transition temperature is
$T_P= 183 $ K.
The zero-temperature
energy gap $\Delta(0)$ is estimated from optical conductivity
data \protect\cite{deg}.
A Fermi velocity of $v_F=2 \times 10^5$ cm/sec was
estimated from band structure calculations \protect\cite{wha}.
$\Delta C$ is the specific heat jump
at the transition \protect\cite{bri} and $\gamma T_P$ is
the normal state electronic specific heat that has been
calculated from the density of states estimated from
magnetic susceptibility measurements \protect\cite{joh}
well above the transition temperature.
The longitudinal coherence length $\xi_{0z}$ has
been estimated from x-ray scattering experiments \protect\cite{gir}.
In both models the dimensionless ratios are independent
of any parameters,
except for $\Delta(0)/k_B T_P$ in
Schulz's model, which is described in Appendix \protect\ref{appsch}.
The rigid lattice theory \protect\cite{all,sch} involves a mean-field
treatment of the single-chain Ginzburg Landau
functional (\protect\ref{aa1}) with the
coefficients (\protect\ref{mfa} - \protect\ref{mfc}).
}
\begin{tabular}{lllcc}
Dimensionless & Experimental& Schulz & Rigid lattice\\*[-0.05in]
ratio & value &model & theory\\
\tableline
$\displaystyle {\Delta(0) \over k_B T_P }$
& $5 \pm 1$ & -- & 1.76\\
$\displaystyle{{\Delta C \over \gamma T_P}}$
& $5 \pm 1$ & 3.4 & 1.43\\
$\displaystyle{{\xi_{z0} T_P \over v_F}}$
& $0.18 \pm 0.04$ & 0.23 & 0.23 \\
\end{tabular}
\label{table1}
\end{table}
\begin{table}
\caption{Parameters for several quasi-one dimensional materials.
The observed transition temperature $T_P$ is always much smaller than
the rigid-lattice transition temperature
$T_{RL}$. The phonons near
$2k_F$, which soften at the transition,
can be treated classically since they have frequencies
of the order of $\Omega(0)$ (estimated from Raman and
neutron scattering) which is much smaller than $T_P$.
The zero-temperature gap $\Delta(0)$, estimated from
the peak in the optical absorption
was used to calculate $T_{RL}$
($ T_{RL}=\Delta(0)/1.76k_B$).}
\begin{tabular}{lcccc}
& $T_P$ (K) & $\Delta(0)$ (meV) & $T_P/T_{RL}$ & $\Omega(0)$ (K)\\
\tableline
K$_{0.3}$MoO$_3$& 183 & 90\tablenotemark[1] & 0.31 & 80
\tablenotemark[2] \\
(TaSe$_4$)$_2$I & 263 & 200$\tablenotemark[3]$ & 0.20 &
130\tablenotemark[4] \\
K$_2$Pt(CN)$_4$Br$_{0.3}$ & 120\tablenotemark[5]& 100\tablenotemark[6]
& 0.18 & 58 \tablenotemark[7]\\
TSeF-TCNQ & 29 & 10\tablenotemark[8] & 0.42 & -- \\
\end{tabular}
\label{table2}
\tablenotetext[1]{ Ref. \cite{deg}}
\tablenotetext[2]{ J. P. Pouget, B. Hennion, C.
Escribe-Filippini, and M. Sato, Phys. Rev. B {\bf 43}, 8421 (1991)}
\tablenotetext[3]{ Ref. \cite{ber}}
\tablenotetext[4]{ S. Sugai, M.Sato, and S. Kurihara, Phys. Rev. B
{\bf 32}, 6809 (1985).}
\tablenotetext[5]{ Complete ordering does not occur \cite{car}.}
\tablenotetext[6]{Ref. \cite{bru}}
\tablenotetext[7]{Ref. \cite{car}}
\tablenotetext[8]{
From activation energy of dc conductivity, Ref. \cite{sco}}
\end{table}
\onecolumn
\widetext
|
\section{Introduction}
\label{sec:intro}
When a normal metal (N) is put in good electrical contact with a
superconductor (S), superconducting order is induced in the normal
metal over distances far greater than any microscopic lengthscale,
either of the normal metal or the superconductor. This induced order
leads to a number of remarkable phenomena, such as the Josephson effect
in SNS junctions \cite{REF:JOSEPH} and the induced Meissner effect in SN
bilayers
\cite{REF:Deutscher}, collectively known as \lq\lq proximity
effects\rq\rq\ \cite{REF:Deutscher}. Until very recently, all work on such
effects, both experimental and theoretical, has concentrated on systems
in which N is in the Fermi-liquid (FL) state. It has long been
appreciated theoretically that, in contrast with their
higher-dimensional analogues, (effectively) one-dimensional systems of
interacting electrons are not Fermi liquids. Instead, they exhibit
a number of possible regimes \cite{REF:fn_states:Emery}, among which
the Luttinger liquid (LL) provides a one-dimensional (1D) metallic
counterpart to the (higher-dimensional) FL state, albeit differing in
several important respects, most notably in the absence of
single-particle excitations in the low-energy part of the spectrum.
The basic features of LLs have been understood mainly in
the context of 1D organic charge-transfer and mixed-valence conductors
\cite{REF:reviews}. In addition, the prediction of
the suppression of the tunneling conductance of LLs \cite{REF:KF1,REF:KF2}
has stimulated the experimental search for
Luttinger liquids in mesoscopic systems, in particular, in the edge-channels of
fractional quantum Hall systems \cite{REF:MILLIK} and in semiconductor
quantum wires \cite{REF:NTT}.
The purpose of this paper is to address the issue of proximity effects
at Luttinger-liquid/super{\-}con{\-}ductor interfaces, including the Josephson
effect in superconductor/Luttinger-liquid/superconductor (SLS)
junction. Our motivation is twofold. First, experimentally, such a
study is relevant in view of the rapid progress in the fabrication of
supercon{\-}ductor/semi{\-}conductor interfaces \cite{REF:exp.review.SSm},
especially those with high inter{\-}face-trans{\-}parency (such as, e.g., the
Nb/InAs interface), and also in view of the recently reported observations
of LL-like behavior in
GaAs quantum wires \cite{REF:NTT}. Thus, the fabrication and
investigation of SLS systems may reasonably be anticipated in the
near future. Second, theoretically, we aim to understand the interplay
between electron-electron interactions and induced superconducting
order in 1D electronic systems. Furthermore, one of the possible
scenarios of high-temperature superconductivity in oxide materials is
built on the assumption of the LL-like character of the normal
electronic state in these materials \cite{REF:ANDERS}. The existence
of LLs in dimensions higher than one, however, is not yet established,
in contrast to the 1D case. Thus, a 1D LL in which the
superconductivity is induced via the proximity effect may provide a model
system for
superconductivity in 2D
\cite{REF:fn_LL_pairing}.
Our main results can be formulated as follows: (i)~At low temperatures,
the Josephson
current through an SLS junction having perfectly transmitting
interfaces has the same phase- and length-dependences as in
the noninteracting case, the only difference being a renormalization
of the effective Fermi velocity. The reason for this is that, using the bosonic
language, the non-dissipative (topological) currents, including the
Josephson current, are carried in LLs by the topological modes of the
boson fields, which are not sensitive to the interactions. At
temperatures above a certain crossover value, interactions lead to the
additional suppression of the Josephson current.
(ii)~The (induced) superconducting condensate wavefunction in a LL in good
electrical contact with a superconductor decays as $x^{-\gamma}$ with
the distance $x$ from the LL/S interface, with $\gamma$ depending on
the strength of the interactions. ($\gamma=1$ corresponds to the
noninteracting case, whereas $\gamma>1$ for repulsive and $\gamma<1$
for attractive interactions.) (iii)~For the case of imperfectly
transmitting interfaces, the renormalization of the interface-transmission
coefficients (via a mechanism known from studies of LLs
coupled by weak links \cite{REF:KF1,REF:KF2}) is reflected in the
renormalization of
the Josephson current, which gets strongly suppressed in the case of
the repulsive interactions. Along the way, we have also: (iv)~derived
effective boundary conditions describing Andreev reflection at the
interface with a superconductor, which we have then used as boundary
conditions in the bosonization procedure; (v)~determined the structure
of the topological (Haldane) excitations in an SLS system; and
(vi)~confirmed result~(i) via an alternative approach, in which
bosonization is applied to both the superconducting and normal parts of
the system.
The issue of the Josephson current through a LL has also been studied
in a recent paper by Fazio, Hekking, and Odintsov \cite{REF:Hekking} for
the case of poor interface-transmittance (see also Ref.~\cite{lutt_and}). By
using the tunneling
Hamiltonian method, it was found that the Josephson current through an
SLS junction is suppressed compared to the noninteracting
case. The present paper takes a different approach. This approach
originates from work on SNS junctions with perfect
interface-transmittance
\cite{REF:Kulik,REF:Ishii,REF:Svid,REF:Bardeen}, in which the Josephson
current was related to the spectrum of electronic states confined to
the N region by Andreev reflection. Our results concerning poorly
transmitting interfaces agree with those of Ref.~\cite{REF:Hekking} (up
to non-universal numerical prefactors, which we have not attempted to
calculate).
The present paper is organized as follows. In Sec.~\ref{sec:abc} we
derive the boundary conditions for the fermion field operators at the
NS interface in the absence of interactions. In Sec.~\ref{sec:bos} we develop
a bosonization
procedure for interacting fermions confined to the normal 1D region of
an SLS system, which makes use of the boundary conditions derived in
Sec.~\ref{sec:abc}. We calculate the Josephson current through an SLS
junction in Sec.~\ref{sec:jos}. In Sec.~\ref{sec:prox}, we analyze the
profile of the condensate amplitude in a semi-infinite LL connected to
a superconductor. Up to this stage, we will have applied bosonization
only to the normal part of the system, the presence of a superconductor
being implemented as a boundary condition. An alternative approach, in
which both the normal and the superconducting parts of the system are
bosonized, is presented in Sec.~\ref{sec:ms}.
\section{Andreev boundary conditions}
\label{sec:abc}
\subsection{Andreev reflection at the NS interface: qualitative picture}
\label{sec:aqp}
Electronic excitations in a normal metal having energies smaller than
the superconducting energy gap $\Delta_0$ suffer Andreev reflection at
the interface, i.e., electron-like excitations are reflected as
hole-like excitations, with a Cooper pair being injected into the
superconductor, and vice versa \cite{REF:Andr,REF:fn.normal}. The
single-particle excitations in S are mixtures of electron- and
hole-like states with weights determined by the self-consistency
condition. In the bulk of N, the electron- and hole-like states are
uncorrelated. Near the boundary, however, Andreev reflection mixes the
electron- and hole-like states precisely in the same proportion as they
are mixed in S, which leads to the formation of a condensate, the
amplitude of which decays into the bulk of N. The decay-length of this
condensate is the length $L_T$ over which superconducting correlations in the
motion of
bulk normal-state electrons exist. In the case of perfect metals,
$L_T=\hbar v_{\rm F}/T$, where $v_{\rm F}$ is the Fermi velocity and $T$ is the
temperature (we choose units in which $k_{\rm B}=1$).
(The same length determines the thermal disruption
(in the absence of inelastic processes) of mesoscopic phase-coherence, as is
manifested
in the phenomenon of universal conductance fluctuations \cite{REF:ucf}).
For $T\ll \Delta\simeq T_{\rm c}$ (where $T_{\rm c}$ is the critical
temperature of
S), $L_T\gg\xi_{\rm S}$ (where $\xi_{\rm S}\simeq\hbar v_{\rm F}/\Delta_0$ is
the coherence length of S).
In order to describe the influence of the superconductors
on the N region, we now derive effective boundary conditions that
account for the Andreev reflection suffered by the low-energy
components of the fermion fields at the NS interfaces. Our
strategy is as follows: in Sec.~\ref{sec:dabc}, we derive these boundary
conditions for
the case of the non-interacting electron gas in N; then, in Sec.~\ref{sec:bos},
we
implement these boundary conditions into the bosonization scheme for
interacting electrons.
\subsection{Derivation of Andreev boundary conditions}
\label{sec:dabc}
We consider a one-dimensional electronic conductor (i.e., a quantum
wire) of length $L$, adiabatically connected to superconducting leads
(see Fig.~\ref{FIG:fig1}a). We begin by analyzing the ideal case, in which the
single-electron parameters (Fermi velocities, effective masses, etc.)
are the same in the N and S parts of the structure, the only difference
between N and S being the presence of a pairing potential in S. We
adopt the conventional model
\cite{REF:Kulik,REF:Ishii,REF:Svid,REF:Bardeen} in which the pairing
potentials in the leads are assumed to be unaffected by the presence of
N. Although this is a non--self-consistent approximation, it is known to
reflect correctly the aspects of the problem relevant for the present
treatment \cite{REF:Kulik,REF:Ishii,REF:Svid,REF:Bardeen}.
We temporarily replace the real 3D superconducting leads by effective
1D leads. The profile of the pair potential is then given by
(cf.~Fig.~\ref{FIG:fig1}b)
\begin{equation}
\Delta(x)=
\cases{
\Delta_0 e^{i\chi_1}&for $x\leq 0$;\cr
0&for $0<x<L$;\cr
\Delta_0e^{i\chi_2}&for $x\geq L$.
}
\label{eq:delta}
\end{equation}
In the Andreev (semi-classical) approximation \cite{REF:Andr},
which is valid
for $\Delta_0\ll\epsilon_F\equiv\hbar k_{\rm F}v_{\rm F}/2$,
the spinor of Bogoliubov amplitudes
\begin{eqnarray}
&&{\bf w}=\left(
\begin{array}{l}
u\\
v\\
\end{array}\right)
\label{eq:nambu}
\end{eqnarray}
satisfying the Bogoliubov-de~Gennes equations \cite{REF:deGennes} is
decomposed into left- and right-moving components,
\begin{equation}
{\bf w}=e^{ik_{\rm F}x}{\bf w}_{+}+e^{-ik_{\rm F}x}{\bf w}_{-},
\label{eq:decomp}\end{equation}
where $k_{\rm F}$ is the Fermi wavevector. The components ${\bf w}_{\pm}$ now
satisfy
the (formally) relativistic (first-order) Bogoliubov-de~Gennes
equations:
$H^{\pm}_D{\bf w}_{\pm}=\epsilon
{\bf w}_{\pm}$, with the
Hamiltonians
\begin{eqnarray}
H^{\pm}_D\: &=& \:
\left(
\begin{array}{cc}
\mp i\hbar v_{\rm F}\partial _x & \Delta(x)\nonumber\\
\Delta^{*}(x) & \pm i\hbar v_{\rm F}\partial _x\nonumber
\end{array}\right).
\label{eq:dirac}
\end{eqnarray}
The full solution of these equations is obtained
\cite{REF:Kulik} by finding the solutions
in the N and S regions and then matching them at the interfaces.
(In the semi-classical approximation, only the wavefunctions
need be continuous.)\thinspace\ The solution in N
for $\epsilon<\Delta_0$
can be written as
\begin{eqnarray}
{\bf w}_{\pm}\: &=& \:
A_{\pm}\left(
\begin{array}{l}
e^{\pm ikx}\\
{\cal R}^{\mp 1}(\epsilon)e^{-i\chi_1}e^{\mp ikx}
\end{array}\right),
\label{eq:bdgp}
\end{eqnarray}
where
\begin{equation}
\label{eq:refl}
{\cal
R}(\epsilon)=e^{-i\eta(\epsilon)}\quad{\rm and}\quad\eta(\epsilon)=
\cos^{-1}(\epsilon/\Delta_0),
\end{equation}
in which ${\cal R}$ is the Andreev reflection coefficient,
whose phase is $\eta$. The quasiparticle momentum
$\hbar k=\epsilon/v_{\rm F}$ satisfies the quantization condition
\begin{equation}
{\cal R}(\epsilon)^2e^{\pm i(\chi_1-\chi_2)}e^{2ikd}=1,
\label{eq:eigen}\end{equation}
where $\pm$ corresponds to two sets of energy levels \cite{REF:Kulik}.
In Eq.~(\ref{eq:bdgp}), $A_{\pm}$ are overall
normalizations which, without the loss of generality, can be chosen to
be real. Evaluating Eq.~(\ref{eq:bdgp}) at $x=0$ and
$x=L$ and using Eq.~(\ref{eq:eigen}), one can see that at the
boundaries the left and right components of the Bogoliubov amplitudes
satisfy
\begin{equation}
\cases{
v_{\pm}={\cal R}^{\mp 1}e^{-i\chi_1}u_{\pm},&for $x=0$,\cr
v_{\pm}={\cal R}^{\pm 1}e^{-i\chi_2}u_{\pm},&for $x=L$.}
\label{eq:uvbc}
\end{equation}
Equations~(\ref{eq:uvbc}) describe the essence of Andreev reflection:
the electron-like excitations ($u_{\pm}$) are converted into hole-like
excitations ($v_{\pm}$), at the same time
acquiring the phase of the order parameter ($\chi_{1,2})$ together
with the phase-shift of the Andreev reflection coefficient ($\eta$).
In the limit $\epsilon\ll \Delta_0$, the phase shift $\eta\to\pi/2$
\cite{fn:eta},
and the boundary conditions (\ref{eq:uvbc}) become energy-independent.
This enables one to derive from Eqs.~(\ref{eq:uvbc}) the boundary
conditions for the real-space fermion operators $\psi_s(x)$,
where $s={\uparrow},{\downarrow}$ denotes the spin projection. These
field operators are related to the $(u,v)$ amplitudes via the
Bogoliubov transformation \cite{REF:deGennes}
\begin{equation}
\psi_{s}(x)=\sum\left(c_su(x)-sc^{\dagger}_{-s}v^{*}_{s}(x)\right),
\label{eq:bt}\end{equation}
where $c_s$ ($c^{\dagger}_s$) is the fermion annihilation (creation)
operator, the sum runs over all
single-particle quantum numbers, and the variable $s$ takes on the
values $+1 (-1)$ for the $\uparrow (\downarrow)$ spin-projections. We
decompose $\psi_s(x)$ into the left- and right-movers:
\begin{equation}
\psi_s(x)=e^{ik_{\rm F}x}\psi_{+,s}(x)+e^{-ik_{\rm F}x}\psi_{-,s}(x).
\label{eq:lrdec}\end{equation}
Substituting decompositions (\ref{eq:decomp}) and (\ref{eq:lrdec}) into
Eq.~(\ref{eq:bt}), we obtain the Bogoliubov transformation for
$\psi_{\pm,s}$:
\begin{equation}
\label{eq:bogol}
\psi_{\pm,s}(x)=\sum
\left(c_su_{\pm}(x)-sc_{-s}^{\dagger}v_{\mp}^{*}(x)\right).
\end{equation}
The boundary conditions for the (Pauli) spinors $\psi_{\pm ,s}$ then
follow upon substitution of Eqs.~(\ref{eq:uvbc}) into
Eq.~(\ref{eq:bogol}), and using $\eta=\pi/2$. After some algebra, we
obtain
\begin{mathletters}
\begin{eqnarray}
\psi_{+,\uparrow}\big\vert_{x=0,L}&=&\mp ie^{i\chi_{1,2}}
\psi_{-,\downarrow}^\dagger\big\vert_{x=0,L}\,\,,\label{eq:lrbc1a}\\
\psi_{+,\downarrow}\big\vert_{x=0,L}&=&\pm ie^{i\chi_{1,2}}
\psi_{-,\uparrow}^{\dagger}\big\vert_{x=0,L}\,\,,
\label{eq:lrbc1b}\end{eqnarray}\end{mathletters}
or, more compactly,
\begin{eqnarray}
\left(
\begin{array}{l}
\psi_{+,\uparrow}
\\
\psi_{+,\downarrow}
\end{array}
\right)\Big\vert_{x=0,L}=\mp ie^{i\chi_{1,2}}\,{\hat T}\left(
\begin{array}{r}
\psi_{-,\uparrow}
\\
\psi_{-,\downarrow}
\end{array}
\right)\Big\vert_{x=0,L}.
\label{eq:lrbc}
\end{eqnarray}
Here, ${\hat T}={\hat g}{\hat C}$ is the time-reversal
operator \cite{REF:deGennes}, with
\begin{eqnarray}
{\hat g}\: &=& \:i{\hat \sigma}_y=
\left(
\begin{array}{cc}
0 & 1\\
-1 & 0
\end{array}\right)
\label{eq:metric}
\end{eqnarray}
and ${\hat C}$ being the Hermitian conjugation operator. The presence of
${\hat T}$ in Eq.~(\ref{eq:lrbc}) signals an important property of
Andreev reflection \cite{REF:Andr}: a reflected excitation is the
time-reversed version of an incident one.
Further insight into the meaning of the boundary conditions
(\ref{eq:lrbc}) can be obtained by employing the chiral symmetry of
left-right fermion fields $\psi_{\pm,s}$. In what follows, we adopt the
methods of Refs.~\cite{REF:Eggert,REF:Polch}, in which the chiral
symmetry of boson fields satisfying Dirichlet or Neumann boundary
conditions was used to derive effective periodic boundary
conditions. The right (left) field describes the propagation of the (formally)
relativistic fermions to the right (left) with the Fermi velocity.
Consequently, in the Heisenberg representation, the space-time
dependence of these fields is given by
\begin{equation}
\psi_{\pm, s}(x,t)=\psi_{\pm, s}(x\mp v_{\rm F}t).
\label{eq:waves}\end{equation}
Using the boundary conditions (\ref{eq:lrbc1a},\ref{eq:lrbc1b}), one
sees that at any instant of time the time-dependent left-moving
fermion fields satisfy
\begin{mathletters}
\begin{eqnarray}
\psi_{-,s}^{\dagger}(v_{\rm F}t)&=&-sie^{-i\chi_1}
\psi_{+,-s}
(-v_{\rm F}t),
\label{eq:bcla}\\
\psi_{-,s}^{\dagger}(L+v_{\rm F}t)&=&s ie^{-i\chi_2}
\psi_{+,-s}(L-v_{\rm F}t).
\label{eq:bclb}
\end{eqnarray}
\end{mathletters}Choosing $t=(L+x)/v_{\rm F}$
in Eq.~(\ref{eq:bclb}), we obtain
$\psi_{-,s}^{\dagger}(x+2L)=sie^{-i\chi_2}
\psi_{+,-s}(-x)$.
Equation~(\ref{eq:bcla}) gives $\psi_{+,s}(-x)
=-sie^{i\chi_1}
\psi_{-,-s}^{\dagger}(x)$ which, in combination with the previous equation,
leads to
\begin{mathletters}
\begin{eqnarray}
&\psi_{-,s}(x+2L,t)=e^{i\pi\vartheta}\psi_{-,s}(x,t),
\label{eq:pera}\\
&\psi_{+,s}(x,t)=-sie^{i\chi_1}\psi_{-,-s}^{\dagger}(-x,t),
\label{eq:perb}\end{eqnarray}
\end{mathletters}where $\vartheta\equiv 1+(\chi/\pi)$ and $\chi\equiv
\chi_2-\chi_1$. Thus, we see that the Andreev boundary
conditions~(\ref{eq:lrbc1a},\ref{eq:lrbc1b}) are equivalent to twisted {\it
periodic}
boundary conditions for $\psi_{-,s}$, Eq.~(\ref{eq:pera}), on an
interval of length twice the length of the original system,
supplemented by the connection between the $\psi_{+,s}$ and
$\psi_{-,s}$ fields following from the chiral symmetry,
Eq.~(\ref{eq:perb}). (Equivalently, the periodic boundary conditions
can be derived for $\psi_{+,s}$ and the chiral symmetry can be used to
obtain $\psi_{-,s}$.)\thinspace\ The problem thus becomes very similar to one
of fermions on a ring of circumference $2L$, threaded by an
effective Aharonov-Bohm flux $\vartheta/2$, the persistent current
\cite{REF:GI,REF:Kulik_pers,REF:BYL} being the analogue of the Josephson
current.
A detailed treatment of persistent currents
in Luttinger liquids was given in Ref.~\cite{REF:Loss} (for the case of
spinless electrons), and we shall adopt this treatment in what follows.
There are significant differences between the system of electrons on a
ring and an SLS junction, however: (i)~On a ring, the number of
electrons is fixed. Therefore, the system is described by the
canonical ensemble, and is sensitive to the parity of the total number
of fermions \cite{REF:Leggett,REF:Loss}. In the normal part of an SLS
junction, the number of electrons
is not fixed, so that the system must be described by the grand
canonical ensemble, the chemical potential being maintained in the
superconducting leads. (ii)~On a ring, all four components of the
fermion fields ($\psi_{\pm,\uparrow,\downarrow}$) satisfy equivalent
twisted periodic boundary conditions. In an SLS junction, only two of
the four components satisfy twisted periodic boundary conditions,
Eq.~(\ref{eq:pera}); once these components are constructed, the other
two are found using the chiral symmetry, Eq.~(\ref{eq:perb}).
We now discuss the range of validity of the Andreev
boundary conditions. The condition $\epsilon\ll \Delta_0$, which we
needed in order to arrive at Eq.~(\ref{eq:lrbc}), means that our
boundary conditions are capable of describing only
excitations with wavelengths $1/k\gg\hbar v_{\rm
F}/\Delta_0\simeq\xi_{\rm S}$. Such excitations exists only in \lq\lq
long\rq\rq\ junctions, i.e., $L\gg\xi_{\rm S}$; thus our treatment is
valid only for this case. On the other hand, as follows from
self-consistent calculations \cite{REF:Deutscher}, the order parameter in
S gets reduced from its bulk value over the scale $\xi_{\rm S}$ near
the boundary, which also affects the excitations in N in the boundary
region of the thickness of $\xi_{\rm S}$. Thus, the model of a
step-like profile of $\Delta$, Eq.~(\ref{eq:delta}), can adequately
describe only processes taking place in the interior of N (i.e., for
$x$ outside boundary layers of width $\xi_{\rm S}$), where the exact shape of
the
profile of $\Delta$ in S is irrelevant. The latter
condition can be satisfied only if
$L\gg\xi_{\rm S}$. Therefore, the range of
validity of our boundary conditions is the same as that of the
non--self-consistent model itself. We can also view Eq.~(\ref{eq:lrbc})
as the minimal-model boundary conditions that describe the
time-reversal process associated with Andreev reflection. Thus, we
now relax the assumption of 1D superconducting leads, and regard
Eqs.~(\ref{eq:lrbc}) as the general boundary conditions satisfied by
the low-energy components of the fermion fields.
\section{Bosonization of Luttinger liquid in contact with superconductors}
\label{sec:bos}
We now turn to the bosonization of an interacting 1D electronic system
\cite{REF:MSB} in contact with superconductor. We represent the free
fermion fields in the conventional bosonic form:
\begin{equation}
\psi_{\pm,s}(x)=\frac{1}{\sqrt{\alpha L}}\exp\left(\pm
i\sqrt{\pi}\phi_{\pm,s}(x)\right),
\label{eq:mand}
\end{equation}
where $\alpha\rightarrow +0$ is a convergence factor and the chiral
bosons $\phi_{\pm,s}$ are expressed through the density (phase) bosons
$\phi_s (\theta_s)$ via
\begin{equation}
\phi_{\pm,s}(x)=\phi_s(x)\mp\theta_s(x).
\label{eq:chbos}\end{equation}
We construct mode expansions for $\phi_{\pm,s}(x)$ in such a way
that the twisted boundary conditions (\ref{eq:pera}) and the auxiliary
conditions (\ref{eq:perb}) are satisfied:
\begin{mathletters}
\begin{eqnarray}
\phi_{-,s}(x)&=&\frac{\varphi_{s}}{\sqrt{\pi}}+\sqrt{\pi}(N_s
+\vartheta)
\frac{x}{2L}+{\bar\phi}_s(x),\label{eq:mmin}\\
\phi_{+,s}(x)&=&\frac{\varphi_{-s}}{\sqrt{\pi}}-\sqrt{\pi}(N_{-s}+\vartheta)
\frac{x}{2L}+{\bar \phi_{-s}(-x)}.\label{eq:mplus}
\end{eqnarray}
\end{mathletters}(The additive c-number terms have been omitted in the
expansion for $\phi_{+,s}$.) Here, $\varphi_s$ are zero-mode operators, $N_s$
are operators whose eigenvalues give the winding numbers of the
Haldane (topological) excitations \cite{REF:Haldane}, and ${\bar
\phi}_s(x)$ are the nonzero-mode components of the chiral boson fields,
which are periodic on the interval $(0,2L)$:
\begin{equation}
{\bar \phi_s(x)}=\sum_{k>0}\gamma_k\Big(e^{-ikx}a^{\dagger}_{k,s}+e^{ikx}
a_{k,s}^{{\phantom\dagger}}\Big),
\label{eq:chirper}
\end{equation}
where $k=\pi n/L$ (with $n=1, 2, \dots$),
$\gamma_k=\exp(-\alpha n/2)/\sqrt{kL}$, and $a_{k,s}$
satisfy the canonical commutation relations
$[a_{k,s}^{\phantom{\dagger}},a^{\dagger}_{k',s'}]=\delta_{ss'}\delta_{kk'}$.
In Eqs.~(\ref{eq:mmin},\ref{eq:mplus}), the terms linear in $x$
describe
the topological excitations of the bosonic system, which do not
conserve the total number of fermions. The eigenvalues of $N_s$ give
the numbers of fermions added to or removed from the Luttinger liquid.
The nonzero-mode components ${\bar \phi_s}$ describe the quantum
fluctuations around the topological excitations. These correspond to
the fluctuations in the fermion density that conserve the total number
of fermions.
We require that the chiral bosons obey the canonical commutation relations
\cite{REF:Shankar}
\begin{equation}
[\phi_{\pm,s}(x),\partial_{x'}\phi_{\pm, s'}(x')]
=\mp i\delta_{ss'}\sum_{n=-\infty}^{\infty}
\delta(x-x'+2nL),
\label{eq:ccrchir}
\end{equation}
where the summation over $n$ reflects periodicity on the interval
(0,2L). The nonzero-mode components of, e.g., expansion
(\ref{eq:mmin}) obey
\begin{eqnarray}
&&[{\bar \phi}_{-,s}(x),\partial_{x'}{\bar \phi}_{-,s'}(x')]\nonumber\\
&&\qquad=\delta_{ss'}\Big\{ix/L+
\sum_{n=-\infty}^{\infty}\delta(x-x'+2nL)\Big\}.
\label{eq:nmccr}\end{eqnarray}
Thus, in order for Eq.~(\ref{eq:ccrchir}) to be satisfied, the
zero-mode operators
$\varphi_s$ and the winding-number operators $N_s$ must obey
\begin{equation}
[\varphi_s,N_{s'}]=2i\delta_{ss'}.
\label{eq:zmnc}\end{equation}
(The same result can certainly be found by considering the commutation
relations of $\phi_{+,s}$.) {\it A posteriori}, we can also justify the
choice of the coefficients $\gamma_k$ in
Eqs.~(\ref{eq:mmin},\ref{eq:mplus}): they were chosen in such a way
that the commutation relations (\ref{eq:ccrchir}) are satisfied.
Next, we introduce the charge ($\rho$) and spin ($\sigma$) components
of the boson fields:
$\phi_{\rho,\sigma}\equiv (\phi_{\uparrow}\pm\phi_{\downarrow})/\sqrt{2}$ and
$\theta_{\rho/\sigma}\equiv
(\theta_{\uparrow}\pm\theta_{\downarrow})/\sqrt{2}$.
The mode expansions for $\phi_{\mu}$ and $\theta_{\mu}$ (where
$\mu=\rho, \sigma$) follow from the expansions
(\ref{eq:mmin},\ref{eq:mplus}):
\begin{mathletters}
\begin{eqnarray}
\phi_{\rho}(x)&=&\frac{\varphi_{\rho}}{\sqrt{\pi}}+\sum_{k>0}\gamma_k\cos kx
\left(a_{k\rho}^{\dagger}+a_{k\rho}^{{\phantom\dagger}}\right),
\label{eq:phirho}\\
\phi_{\sigma}(x)&=&\sqrt{\frac{\pi}{2}}M\frac{x}{L}+i\sum_{k>0}\gamma_k
\sin kx \left(a_{k\sigma}^{\dagger}-a_{k\sigma}^{{\phantom\dagger}}\right),
\label{eq:phisig}\\
\theta_{\rho}(x)&=&\sqrt{\frac{\pi}{2}}(J+\vartheta)\frac{x}{L}
+i\sum_{k>0}\gamma_k\sin kx
\left(a_{k\rho}^{\dagger}-a_{k\rho}^{{\phantom\dagger}}\right),
\label{eq:thetarho}\\
\theta_{\sigma}(x)&=&\frac{\varphi_{\sigma}}{\sqrt{\pi}}
+\sum_{k>0}\gamma_k\cos kx
\left(a_{k\sigma}^{\dagger}+a_{k\sigma}^{{\phantom\dagger}}\right),
\label{eq:thetasig}
\end{eqnarray}
\end{mathletters}where $\varphi_{\rho/\sigma}
\equiv(\varphi_{\uparrow}\pm\varphi_{\downarrow})/
\sqrt{2}$, $M\equiv (N_{\uparrow}-N_{\downarrow})/2$,
$J\equiv (N_{\uparrow}+N_{\downarrow})/2$, and
$a_{k\rho/\sigma}\equiv(a_{k,\uparrow}\pm a_{k,\downarrow})/\sqrt{2}$.
It is natural that the phase difference of the superconducting
order parameters $\chi$,
which determines the charge flow between the superconductors, appears
only in the field associated with the charge
current, i.e., $\theta_{\rho}$.
We now have to determine the topological constraints imposed on the winding
numbers $N_s$ (and, consequently, on $M$ and $J$). This can be done by
substituting the expansion, e.g., for $\phi_{-,s}$, Eq.~(\ref{eq:mmin}),
into the bosonization formula (\ref{eq:mand}), and requiring that
the boundary conditions for fermions (\ref{eq:pera}) be
satisfied \cite{REF:Loss}. [When disentangling the operators in the
exponent of Eq.~(\ref{eq:mand}), one must recall that $\phi_s$ and $N_s$
do not commute, and use Eq.~(\ref{eq:zmnc}).]\thinspace\
One thus finds that $N_s$ satisfies
\begin{equation}
(-)^{N_s+1}=1,
\label{eq:constr}\end{equation}
i.e., that the eigenvalues of $N_s$ are odd.
(Neglecting the operator-nature of the zero-modes
and the winding numbers would have led to $N_s$ being even.) Consequently,
$J+M$ must be odd. It is convenient to introduce an effective winding-number
$J'=J+1$, so that $J+\vartheta=J'+\chi/\pi$ in Eq.~(\ref{eq:thetarho}). Then,
$J'+M$ must be even. Comparing this
constraint with the similar constraint on the topological numbers in the
persistent-current problem \cite{REF:Loss}, we see that our constraint
effectively
corresponds to the case of an {\it odd\/} number of fermions on the
ring, in which case the response of the system to the twist in boundary
conditions
is diamagnetic, i.e., the free energy is minimal at zero twist. Tracing back
through our calculations, we note that the diamagnetic nature of the
Josephson current is guaranteed by the Andreev phase shift ($\pi/2$),
which ultimately shifts $J$ to $J+1$.
The physical meaning of the topological constraint is quite simple:
the energy of the LL is minimal when the left- and right-moving branches
of the spectrum are populated symmetrically; changing the total number of
fermions in the LL by an even (odd)
number results in excitations
with even (odd) total momentum quanta.
We note that expansions similar to
Eqs.~(\ref{eq:phirho},\ref{eq:phisig},\ref{eq:thetarho},\ref{eq:thetasig})
could have been
obtained by first deriving the boundary conditions directly
for the charge and spin bosons
from the boundary conditions for the fermions (in the same way that the
boundary
conditions for bosons are derived from the Dirichlet boundary conditions for
fermions
in Ref.~\cite{REF:Eggert}), and then constructing expansions satisfying
these boundary conditions. In this way, however, the zero-modes of the
expansions, which are crucial for the topological constraints on the
eigenvalues of $M$ and
$J$, might have
been missed. (We will derive and use the boundary conditions for charge/spin
bosons
in Sec.~\ref{sec:prox}, when the
topological structure of the boson fields will not be important.)\thinspace\
The bosonized Hamiltonian of the LL is given by
\begin{equation}
{\cal H}=\frac{\hbar}{2}\!\sum_{\mu=\rho,\sigma}\!\int_{-L/2}^{L/2}dx\left\{
\frac{v_{\mu}}{K_{\mu}}(\partial_x \phi_\mu)^2+
v_{\mu}K_{\mu}(\partial_x \theta_\mu)^2\right\}.
\label{eq:llham}\end{equation}
If the LL model originates from the Hubbard model then
$K_{\rho/\sigma}=1/\sqrt{1\pm g}$, where $g\equiv Ua/\pi v_{F}$,
with $U$ being the strength of on-site interactions and
$a$ the microscopic length cut-off (of order
the Fermi wavelength), and $v_{\mu}\equiv
v_{F}/K_{\mu}$. In addition, if the underlying SU(2) symmetry of the
Hubbard model is intact, then $K_\sigma=1$ \cite{REF:Emery}.
\section{Josephson current through a Luttinger liquid}
\label{sec:jos}
One of the most important consequences of induced coherence in the
N part of an SNS system, is the
Josephson current through it. This current differs from its counterpart
in tunnel junctions in that the critical current ${\cal J}_c$ decays
with the junction length $L$ according to the power-law $1/L$
(for $L\ll L_T$), rather than
exponentially.
The Josephson current in SNS junctions is affected strongly by the
quality of the interface. The transmittance of interfaces between
semiconductors and superconductors varies widely, depending on the
nature of the junction. The interface that has been studied most
intensively in recent years, particularly in the context of mesoscopic
effects, is the Nb/InAs interface. This interface is unique in the
sense that a charge-accumulation layer is formed instead of a Schottky
barrier and, as a result, the interface transparency is quite high.
(According to a recent measurement of the proximity effect in this
structure, the interface transmission coefficient $T_0\approx 0.7$
\cite{REF:vanwees}.)\thinspace\ More commonly, however, the
transmittance may be quite low, both because of interface roughness and
Schottky-barriers formation. Below, we calculate the Josephson current
through the LL in two limiting cases: perfectly transmitting interfaces
(Sec.~\ref{sec:hti}) and poorly transmitting interfaces
(Sec.~\ref{sec:pti}). The latter case has been investigated in
Ref.~\cite{REF:Hekking}.
\subsection{Perfectly transmitting interfaces}
\label{sec:hti}
First, we consider the case of perfectly transmitting interfaces,
in which the only scattering that takes place at the S/LL boundaries is
Andreev reflection, single-particle reflection being absent. The
Josephson current ${\cal J}$ is given by
\begin{equation}
\label{eq:jcdef}
{\cal J}=\frac{2e}{\hbar}\frac{\partial}{\partial \chi}\Omega,
\end{equation}
where $\Omega=-k_{B}T\ln Z$ is the grand po{\-}tential, and $Z$ is the grand
partition function.
Substituting
Eqs.~(\ref{eq:phirho},\ref{eq:phisig},\ref{eq:thetarho},\ref{eq:thetasig})
into Eq.~(\ref{eq:llham})
and diagonalizing the nonzero-mode part via a canonical transformation,
we get for the many-body eigenenergies
of the system
\begin{eqnarray}
{\cal E}&=&\frac{\pi \hbar}{4L}\left[v_{\rho}K_{\rho}\left(J'+
\frac{\chi}{\pi}\right)^2+\frac{v_{\sigma}}{K_{\sigma}}
M^2\right]
\nonumber\\
&&\qquad\qquad+\hbar \sum _{k>0}\sum_{\mu=\rho,\sigma}
v_{\mu}k(n_{k \mu}+1/2),
\label{eq:eigenst}\end{eqnarray}
where $n_{k \mu}^{}\equiv b_{k \mu}^{\dagger}b_{k \mu}^{}$, and
the new boson operators $b_{k \mu}$ are connected to
the old ones via
$a_{\mu}=b_{\mu}\cosh\lambda_{\mu} -\nolinebreak
b_{\mu}^{\dagger}\sinh\lambda_{\mu}$, in which
\begin{equation}
\label{eq:canon}
\lambda_{\rho/\sigma}=\pm\frac{1}{2}\tanh^{-1}\frac{1-K_{\rho/\sigma}^2}
{1+K_{\rho/\sigma}^2}.
\end{equation}
We see that the phase-difference $\chi$ appears only in the topological part
of ${\cal E}$, as it should, because the nonzero-mode excitations are neutral,
and therefore do not contribute to the (equilibrium) charge current.
We also note that only two of the four
charge/spin bosons, viz., $\phi_{\sigma}$ and $\theta_{\rho}$,
contribute to the topological part of ${\cal E}$. (The Josephson-current
problem differs in this respect from the persistent-current problem, in
which all four bosons contain topological excitations.)\thinspace\
The combination $\{\phi_{\sigma},\theta_{\rho}\}$ commonly arises in the study
of
superconductivity in LLs \cite{REF:Emery}.
The partition function factorizes as $Z=Z_{\rm t}(\chi)Z_{\rm n}$,
where $Z_{\rm t/n}$ is the contribution from the topological
(nonzero-mode) part of ${\cal E}$. To calculate ${\cal J}$,
we need only know $Z_{\rm t}$, which is given by
\begin{equation}
\label{eq:zt1}
Z_{\rm t}(\chi)=
{\sum\limits_{J',M}}^{\prime}
e^{-\varepsilon_{\rho}
\left(J'+\chi/\pi\right)^2} e^{-\varepsilon_{\sigma}M^2},
\end{equation}
where $\varepsilon_{\rho}\equiv\pi L_Tv_{\rho}K_{\rho}/4v_{\rm F}L$
and $\varepsilon_{\sigma}\equiv\pi v_{\sigma}L_T/4K_{\sigma}v_{\rm F}L$,
and the primed sum indicates that $J'$ and $M$ are connected via the
constraint
found in Sec.~\ref{sec:bos} (i.e., $J'+M$ even).
(Although the spin-part of $Z_{\rm t}$ does not depend on $\chi$,
it does not simply reduce to an overall multiplicative factor because of this
constraint.)\thinspace\ It is convenient to represent the winding numbers
$J'$ and $M$ in the following form:
$J'=2j+\kappa_{\rm J}$ and
$M=2m+\kappa_{\rm M}$ (with $\kappa_{\rm J}=0,1$ and $\kappa_{\rm M}=0,1$)
\cite{REF:Loss}.
The topological constraint is then satisfied
for $j,m=0,\pm 1,\dots$ and
$\kappa_{\rm J}=\kappa_{\rm M}$.
We can then re-write Eq.~(\ref{eq:zt1}) in the unconstrained form:
\begin{equation}
\label{eq:zt2}
Z_{\rm t}(\chi)=
f_{0,\rho}(\chi)f_{0,\sigma}(0)+f_{1,\rho}(\chi)f_{1,\sigma}(0),
\end{equation}
where
\begin{equation}
\label{eq:sums}
f_{\kappa,\mu}(\chi)\equiv
\sum_{n=-\infty}^{\infty}e^{-\varepsilon_{\mu}\left(2n
+\kappa+\chi\right)^2}.
\end{equation}
The Josephson current can now readily be calculated. Without writing
down the exact expression (which contains, as usual for this kind of
problem, Jacobi $\vartheta_{3}$-functions \cite{REF:Loss}), we consider
only the asymptotic cases of low ($L\ll L_T$) and high
($L\gg L_T$) temperatures. In the former case, one finds
\begin{equation}
\label{eq:jlow}
{\cal J}=\frac{ev_{\rho}K_{\rho}}{L}\;\;\frac{\chi}{\pi},\;\;
\mbox{for}\;\;|\chi|\leq\pi,
\end{equation}
with ${\cal J}(\chi+2\pi)={\cal J}(\chi)$.
We note that the interaction-renormalization
of the Josephson current is the same
as that of the persistent current \cite{REF:Loss}.
When the Luttinger-liquid Hamiltonian (\ref{eq:llham}) is
obtained as the long-wavelength limit of the Hubbard Hamiltonian then
$v_{\rho}K_{\rho}=v_{\rm F}$. Comparing Eq.~(\ref{eq:jlow}) for
$v_{\rho}K_{\rho}=v_{\rm F}$ with the
corresponding expressions for the non-interacting electrons
\cite{REF:Ishii,REF:Svid,REF:Bardeen}, we see that {\it the Josephson
current through the Luttinger liquid is precisely the same as through
the non-interacting electron gas}. A word of caution is necessary,
however: this conclusion is only valid if backscattering and Umklapp
scattering are not taken into account. Even if these types of
scattering are irrelevant (in the renormalization-group sense), they
will modify the parameters of the LL entering Eq.~(\ref{eq:llham}),
so that the equality $v_{\rho}K_{\rho}=v_{\rm F}$ will no longer hold
\cite{fn_irr}.
Nevertheless, the deviations from this equality are expected to
be small. (For instance, in the spinless case, the maximal
reduction in the product $v_{\rho}K_{\rho}$ due to Umklapp scattering
is 20\%, even at half-filling, when such processes are most effective
\cite{REF:Loss,REF:LM}).)\thinspace\ Also, there is a much more
significant source of the renormalization of ${\cal J}$ which we have
not yet taken into account, viz., the non-perfectness of the interfaces
(see Sec.~\ref{sec:pti}).
For high temperatures ($L\gg L_T$) we find
\begin{equation}
\label{eq:jhigh}
{\cal J}=8e\hbar^{-1}T
\exp\left(-2\pi\alpha L/L_T
\right)\sin\chi,
\end{equation}
where
\begin{equation}
\alpha\equiv
\frac{1}{2}\left(\frac{v_{\rho}K_{\rho}}{v_{\rm F}}
+\frac{v_{\sigma}}{v_{\rm F}K_{\sigma}}\right).
\label{eq:alpha}\end{equation}
In the Hubbard model with the SU(2) symmetry,
$\alpha=(1+1/\sqrt{1-g})/2>1$. Thus, at high
temperatures, interactions lead to the further suppression of ${\cal J}$
(in addition to the thermal disruption of the phase-coherence).
\subsection{Poorly transmitting interfaces}
\label{sec:pti}
Having discussed the case of perfectly transmitting interfaces, we now
give a brief discussion of the case of poorly-transmitting interfaces.
In this case, qualitative information can be obtained by making use of the
known results on the interaction-induced renormalizations of the
transmission coefficient. (For the analogous treatment of persistent
current in imperfect LL rings, cf.~Ref.~\cite{REF:GP}.)\thinspace\
First, consider a non-interacting SNS \lq\lq clean\rq\rq\ system (i.e.,
the elastic
mean free path being far greater than L), with interface transmission
coefficients $T_{1,2}\ll 1$. For simplicity, we now restrict attention
to the low-temperature case ($L_T\gg L$). The result for
the Josephson current can be obtained from the general formula of
Ref.~\cite{REF:ALO}, Eq.~(16), which expresses $\cal J$ through the
probability for an excitation to propagate from one interface to
another within a certain time. Substituting the probability of
ballistic propagation into Eq.~(16) of Ref.~\cite{REF:ALO}, we find,
after some simple algebra, that the critical current ${\cal J}_c^{i}$
for the structure with imperfect interfaces is
\begin{equation}
{\cal J}_c^{i}\simeq T_1 T_2{\cal J}_c,
\label{eq:jcimp}\end{equation}
where ${\cal J}_c$ is the critical current for $T_1=T_2=1$, which is given
by the $\chi$-independent factor in Eq.~(\ref{eq:jlow}). In the
interacting case, the low-transparency interfaces can be described
within the weak-link model of Kane and Fisher \cite{REF:KF1,REF:KF2}. In this
model, the weak links (in our case, the interfaces) are treated as
perturbations that transfer electrons between the disconnected (in zeroth
order) parts of the LL, and the renormalization group (RG) of the
boundary sine-Gordon model is used to find the effective values of the
hopping (transmission) probabilities. The case of the double barrier
has been considered in Ref.~\cite{REF:KF2}. We note that: (i)~because in our
Hamiltonian (\ref{eq:llham}), and hence in our action, the topological
excitations are decoupled from the nonzero-modes, the RG flow equations
for the transmission coefficients are the same as in Ref.~\cite{REF:KF2};
(ii)~the effective cut-off for the RG flow is provided in our case by
the junction length $L$. Therefore, we can borrow the result for the
renormalized product $T_1 T_2$ from the Kane-Fisher result for the
double barrier away from the resonance:
\begin{equation}
T_1T_2\rightarrow
T_1T_2(k_{\rm F}L)^{-\left(1/K_\rho+1/K_\sigma-2\right)}.
\label{eq:theft}\end{equation}
If the SU(2)
symmetry of the underlying Hubbard model is intact, i.e., $K_\sigma=1$,
we find
\begin{equation}
\label{eq:jcimpint}
{\cal J}_c^{i}
\simeq
T_1 T_2\left(\frac{1}{k_{\rm F}L}\right)^{K_\rho^{-1}-1}
{\cal J}_c,
\end{equation}
which is in agreement with Ref.~\cite{REF:Hekking},
up to a (non--universal) numerical coefficient.
Whether Eq.~(\ref{eq:jlow}) or Eq.~(\ref{eq:jcimpint}) is relevant to a
given experimental situation, depends on the bare (i.e.,
unrenormalized) values of $T_{1,2}$ and on $L$. Suppose that
$T_{1}\approx T_{2}\equiv T_0\approx 1$. Then the interface barriers
can be treated according to the weak-barrier model
\cite{REF:KF1,REF:KF2}. Assume, for simplicity, that the potential
barriers are $\delta$-functions with the (bare) amplitude $V_{0}$. As $V_{0}$
is small, its RG flow at distinct interfaces is independent, and given
by $V=V_0(L/a)^{(1-K_{\rho})/2}$. Then $T_0$ is renormalized to
\begin{equation}
T=\frac{1}{1+(mV/\hbar
k_{\rm F})^2}=\frac{1}{1+\frac{1-T_0}{T_0}
\left(\frac{L}{a}\right)^{(1-K_{\rho})}}.
\end{equation}
For relatively short junctions, i.e.,
\begin{equation}
L\ll L^{*}
\simeq a\left(\frac{T_0}{1-T_0}\right)^{1/(1-K_\rho)},
\label{eq:dstar}
\end{equation}
the renormalization of $T_0$ due to interactions is small,
and Eq.~(\ref{eq:jlow}) applies. The better the interface the
larger $L^*$. In particular, as $T_0\rightarrow 1$, $L^*\rightarrow\infty$,
in accordance with the previously-found virtual absence of the
renormalization of $\cal J$ for perfect interfaces (cf.~Sec.~\ref{sec:hti}).
For longer junctions, i.e., $L\gg L^*$, Eq.~(\ref{eq:jcimpint}) applies.
Choosing $g=Ua/\pi v_{\rm F}=1/2$ (i.e., $K_{\rho}\approx 0.8$), using
the value $T_{0}=0.7$ \cite{REF:vanwees} and recalling
that $a\simeq 1/k_{\rm F}\simeq 100$\AA\ in the relevant semiconductor
structures,
we find
$L^*\simeq 1\,\mu$m. Junctions of lengths in the range $0.1-10\,\mu$m are
quite common in experiments \cite{REF:exp.review.SSm}, so both
Eqs.~(\ref{eq:jlow}) and
(\ref{eq:jcimpint}) may be relevant in experimental situation.
Indirectly, one also can appreciate the extent to which interactions
renormalize the Josephson current by using recent experimental results on
the (dissipative) conductance of the
ultra-high mobility GaAs quantum wires \cite{REF:NTT}. As was shown
in Ref.~\cite{REF:NTT}, the conductance of the wire is reduced from
the conductance quantum (i.e., $e^2/h$ per spin projection) as the temperature
is lowered, the temperature-dependence being consistent with the theory
of charge transport in dirty Luttinger liquids
\cite{REF:Apel,REF:Fukuyama,REF:recent}.
The absolute value of this reduction is quite small, however: it amounts
to $1-5\%$ for wires of length $2-10\,\mu$m.
\section{Proximity effect in Luttinger liquids}
\label{sec:prox}
As has been mentioned in Sec.~\ref{sec:aqp}, Andreev
reflection at the NS interface gives rise to
correlations
between electron- and hole-like excitations in N.
These correlations are similar to those between the single-particle
excitations in S, which can be viewed as the induction of superconducting
off-diagonal long-range order in N due to the proximity of S.
The presence of such order is usually
described by the (inhomogeneous) condensate wavefunction \cite{REF:Deutscher}
$F(x)$, defined by
\begin{equation}
F(x)\equiv\langle
\psi_{\uparrow}(x)\psi_{\downarrow}(x)\rangle
\label{eq:fdef}.
\end{equation}
In the bulk of N, $F=0$ . The scale over which $F$ (exponentially) changes
from its value
at the NS boundary to zero in the bulk is given by $L_T$. As $T\to 0$, the
length $L_T\to\infty$, and the exponential decay of $F$ crosses over to
a slower
(power-law) decay. In particular, if N is a Fermi-liquid metal, $F$ decays with
the distance from the interface as $1/x$ (at $T=0$) \cite{REF:Falk}.
We now explore
how this decay law is changed if N is in the LL state.
Consider a semi-infinite LL occupying the half-line $x>\nolinebreak 0$ and
connected to
S at $x=0$.
The bosonized form of $F(x)$ is given by
\begin{equation}
\label{eq:Fbos}
F(x)=\frac{1}
{\pi\delta a}\Big\langle e^{-i\sqrt{2\pi}\theta_{\rho}(x,0)}
\cos\big(\sqrt{2\pi}
\phi_{\sigma}(x,0)\big)
\Big\rangle,\end{equation}
where $\delta\to +0$ is a (dimensionless) cut-off parameter, $a$ is the
microscopic scale of the system, and $\theta_{\rho}(x,\tau)$
and $\phi_{\sigma}(x,\tau)$ are boson fields in the (imaginary time)
Heisenberg representation. In Eq.~(\ref{eq:Fbos}),
the average is taken with respect to Boltzmann factor $e^{-S/\hbar}$, where
$S=S_{\rho}+S_{\sigma}$
is the (Euclidean) action corresponding to Hamiltonian Eq.~(\ref{eq:llham}),
and
\begin{mathletters}
\begin{eqnarray}
S_{\rho} &\equiv& \frac{\hbar K_{\rho}}{2}\int dx d\tau \frac{1}{v_{\rho}}\left
(\partial_\tau\theta_{\rho}\right)^2+v_{\rho}
\left
(\partial_x\theta_{\rho}\right)^2,\label{eq:action_c}\\
S_{\sigma} &\equiv& \frac{\hbar}{2K_{\sigma}}
\int dx d\tau \frac{1}{v_{\sigma}}\left
(\partial_{\tau}\phi_{\sigma}\right)^2+v_{\sigma}
\left
(\partial_{x}\phi_{\sigma}\right)^2.
\label{eq:action_s}\end{eqnarray}\end{mathletters}(Note
that we have deliberately expressed $S_{\mu}$ via those boson fields
that enter the bosonized form of $F(x)$.)\thinspace\ The presence of S at $x=0$
imposes certain
boundary conditions on these fields. We derive these boundary conditions
directly from the boundary conditions for fermions
Eq.~(\ref{eq:lrbc1a},\ref{eq:lrbc1b}) by using the bosonized form of the
fermion fields (\ref{eq:mand}). (The phase of the order parameter
in S is now taken to be zero,
as we do not consider charge-flow through the
the interface.)\thinspace\ Simple algebra then leads to:
\begin{equation}
\label{eq:bcinf}
\phi_{\sigma}(0,\tau)=-\sqrt{2\pi}/4,\qquad \theta_{\rho}(0,\tau)=0.
\end{equation}
In the (semi)-infinite geometry, the energy of
topological excitations is infinitesimally small \cite{REF:fn.topol.en.}, and
therefore we do not have to
incorporate the winding numbers of such excitations in boundary conditions
(\ref{eq:bcinf}). As one might have anticipated, Andreev boundary conditions
for fermions (\ref{eq:lrbc1a},\ref{eq:lrbc1b}) impose boundary conditions only
on those components of the boson fields that occur in the bosonized form
of the condensate wavefunction (\ref{eq:Fbos}).
In order to remove the divergence in Eq.~(\ref{eq:Fbos})
as $\delta\rightarrow +0$, we
use the following trick. Consider the modified boundary
condition for the $\phi_{\sigma}$-field:
$\phi_{\sigma}(0,\tau)=-\sqrt{2\pi}/4+\delta$. Introduce a new field
${\tilde\phi}_{\sigma}\equiv\sqrt{2\pi}/4-\delta+\phi_{\sigma}$ satisfying
the homogeneous boundary condition ${\tilde\phi}_{\sigma}(0,\tau)=0$.
After this, $F$ takes the form
\begin{eqnarray}
\label{eq:F.interm}
F(x) &=& \frac{\sin\delta}{\delta}\frac{1}{\pi a}
\Big\langle e^{-\sqrt{2\pi}\theta_\rho}\Big\rangle
\Big\langle
e^{-\sqrt{2\pi}{\tilde\phi}_\sigma}\Big\rangle
\Big|_{\delta\rightarrow 0}\nonumber\\
&=&
\frac{1}{\pi a}\exp\left\{-\pi\left(
G_{\rho}(x,x,0)+G_{\sigma}(x,x,0)\right)\right\},
\end{eqnarray}
where $G_{\rho/\sigma}(x,x',\tau)$ is the propagator of the charge
(spin) boson field, which satisfies
\begin{equation}
\label{eq:laplace}
K_{\mu}^{\alpha_{\mu}}(\partial_x^2+v_{\mu}^{-2}\partial_\tau^2)G_{\mu}=
-\delta(x-x')\delta(\tau),
\end{equation}
in which $\alpha_{\rho/\sigma}=\pm 1$. $G_\mu$ obeys the following boundary
conditions: $G_\mu (0,x',\tau)=0$, $G_{\mu}(x,x',\tau)|_{x\to \infty}
\rightarrow 0$ and $G_{\mu}(x,x',\tau+\beta)=G_{\mu}(x,x',\tau)$, where
$\beta=1/T$. The Fourier transform in $\tau$ of the solution of
Eq.~(\ref{eq:laplace}) is given by
\begin{equation}
\label{eq:G}
G_{\mu}(x,x',\omega)=K_{\mu}^{-\alpha_\mu}|{\bar \omega}|^{-1}
\sinh (|{\bar \omega}|x_{<})\exp(-|{\bar \omega}|x_{>}),
\end{equation}
where $x_{<}\equiv\mbox{min}\{x,x'\}$ and $x_{>}\equiv\mbox{max}\{x,x'\}$.
Inverting the transform, we get
\begin{equation}
\label{eq:asympt}
G_{\mu}(x,x,0)=\frac{1}{2\pi K^{\alpha_{\mu}}}\ln\left(x/a\right),
\end{equation}
where, in order to regularize $G_{\mu}$,
we have chosen the same short-distance cut-off $a$
as in Eq.~(\ref{eq:Fbos}).
Substituting Eq.~(\ref{eq:asympt}) into Eq.~(\ref{eq:F.interm}), we find
\begin{equation}
\label{eq:scaling}
F(x)=\frac{C}{a}\left(\frac{a}{x}\right)^{\gamma},\;\;\mbox{with}\;\;
\gamma\equiv\frac{1}{2}(K_{\sigma}+K_{\rho}^{-1}),
\end{equation}
where $C$ is a (non-universal) numerical coefficient.
In the absence of interactions, $K_{\rho}=K_{\sigma}=1$ and we return
to the $1/x$ scaling. In the presence of repulsive (attractive)
interactions, $\gamma>1$ ($\gamma<1$), and
the condensate amplitude in the LL decays faster (slower) than in the FL.
This result is in accord with one's intuition: the repulsive (attractive)
Coulomb interaction
weakens (strengthens) the superconducting
state induced in N by Andreev reflection.
The exponent $\gamma$ is one half of the
exponent determining the spatial decay
of the (singlet) superconducting fluctuations
in the infinite LL \cite{REF:Emery}.
At the first sight, the result that the profile of the condensate wavefunction
in the LL decays faster than in the FL seems to contradict
to the results of Sec.~\ref{sec:jos}, in which it was found
that the junction-length
dependence of the critical current is the same in the LL and the FL. Indeed, it
seems
natural to connect the $1/x$ decay law of the condensate in the FL with
the $1/L$ dependence of ${\cal J}_c$; then, it would be reasonable to expect
that the
$1/x^\gamma$ decay law of the condensate in the LL would be
transformed into a $1/L^{\gamma}$-dependence of ${\cal J}_c$, even if
the interfaces are perfect \cite{fn_gamma}
In fact, this conclusion
would not be valid and, as we show below, the $1/L$-dependence of ${\cal J}_c$
(at $T=0$)
is universal, and not connected with the profile of the
condensate in the N region of an SNS junction. Consider, again, an SNS junction
of length $L\gg \xi_{\rm S}$. Our main argument is that at $T=0$
the only relevant lengthscale in the problem is $L$; therefore, at
distances from the interface larger than $\xi_{\rm S}$, the profile
of condensate
wavefunction N is described by a single dimensionless parameter
$x/L$. Therefore, $F(x)$ can be represented in the following form:
\begin{equation}
F(x)=F_0^{-}\Phi^{-}(x/L)+F_0^{+}\Phi^{+}(x/L),
\label{eq:single}
\end{equation}
where $F_{0}^{\pm}=F_0e^{i\chi_{1,2}}$ are the values of $F$ at
$x=0(L)$, and the scaling functions $\Phi^{\pm}(z)$
satisfy the following boundary conditions: $\Phi^{\pm}(1)=1 (0)$,
$\Phi^{\pm}(0)=0(1)$. The supercurrent flowing through the junction
is given by
\begin{equation}
\label{eq:current}
{\cal J}=iA\left(F(x)\frac{dF^{*}(x)}{dx}-\mbox{c.c.}\right),
\end{equation}
where $A$ is an $L$-independent constant. Substituting Eq.~(\ref{eq:single})
into Eq.~(\ref{eq:current}), we see that ${\cal J}$ can be represented in the
form
\begin{equation}
\label{eq:1/d}
{\cal J}=AL^{-1}\Phi(x/L),
\end{equation}
where $\Phi(z)$ is a scaling function that is a combination
of the functions $\Phi^{\pm}(z)$ and their derivatives.
Due to charge conservation,
${\cal J}(x)$ does not depend on $x$, which can be satisfied only
if $\Phi(z)$ is $z$-independent.
Thus, we see that
$J\propto 1/L$ regardless of
the particular form of the condensate wavefunction,
the latter determining only the numerical coefficient in front
of the $1/L$-dependence.
\section{Solution via the bosonization of the whole system}\
\label{sec:ms}
So far, we have applied bosonization only to the Luttinger-liquid part
of the system, i.e., to the interval $0<x<L$. We can gain some
further insight into our results by comparison with a system in which the
LL occupies the entire real line, but, by some mechanism,
has acquired a superconducting gap when $x<0$ and $x>L$
(cf.~Fig.~\ref{FIG:fig1}c). The
existence of a gap means that the usual Luttinger Hamiltonian is
modified by the addition of the term
\begin{equation}
H_{\rm gap}=\int dx |\Delta(x)|e^{i\chi(x)}
\left(
\psi^{\dagger}_{+\uparrow}\psi^{\dagger}_{-\downarrow}
+\psi^{\dagger}_{-\uparrow}\psi^{\dagger}_{+\downarrow}
\right)+\mbox{H.c}.
\label{EQ:ms_Xone}
\end{equation}
The corresponding (Minkowski) bosonic action is
\begin{eqnarray}
S
&=&
\int dx dt
\Bigl\{
\frac{K_\rho}{2}
\left(
\frac 1{v_\rho}(\partial_t \theta_\rho)^2
-v_\rho(\partial_x \theta_\rho)^2
\right)
\nonumber
\\
&&
\qquad\qquad
+
\frac 1{2K_\sigma}
\left(
\frac 1{v_\sigma}(\partial_t
\phi_\sigma)^2-
v_\sigma(\partial_x \phi_\sigma)^2
\right)
\nonumber
\\
&&
\qquad\qquad
+
|\Delta|
\Big(
:\cos(\chi(x)+\sqrt{2\pi}
(\theta_\rho-\phi_\sigma)):
\nonumber
\\
&&
\qquad\qquad
+
:\cos(\chi(x)+\sqrt{2\pi}(\theta_\rho+\phi_\sigma)):
\Big)
\Bigr\},
\label{EQ:ms_Xtwo}
\end{eqnarray}where $\chi(x)$ is the local phase of the order parameter.
In regions where $\Delta(x)$ is large, the principal effect of the
non-linear terms is to constrain the values of $\theta_\rho(x)$ and
$\varphi_\sigma(x)$ to the minima of the cosine potential, so that
\begin{mathletters}
\begin{eqnarray}
\chi+\sqrt{2\pi}(\theta_\rho-\phi_\sigma)
&=&
2n\pi;
\label{EQ:ms_XthreeA}
\\
\chi+\sqrt{2\pi}(\theta_\rho+\phi_\sigma)
&=&
2m\pi.
\label{EQ:ms_XthreeB}
\end{eqnarray}
\end{mathletters}Equivalently
\begin{mathletters}
\begin{eqnarray}
\theta_\rho
&=&
\frac{1}{\sqrt{2\pi}}(-\chi+\pi(n+m));
\label{EQ:ms_XfourA}
\\
\phi_\sigma
&=&
\frac{1}{\sqrt{2\pi}}(+\pi(n-m)).
\label{EQ:ms_XfourB}
\end{eqnarray}
\end{mathletters}There are no constraints
on $\theta_\sigma$ or $\phi_\rho$. The fields $\theta_{\rho}$ and
$\phi_{\sigma}$
are thus locked (modulo winding numbers) to the condensate phase in the
superconducting regions. In the purely LL part of the
system all four fields are free to fluctuate. The condensate therefore
imposes boundary conditions that are essentially the same as those in
Eq.~(\ref{eq:bcinf}).
The bosonized form of the number density current is
\begin{equation}
j(x)=-2v_\rho K_\rho\frac
1{\sqrt{2\pi}}\partial_x\theta_\rho.
\label{EQ:ms_Xfive}
\end{equation}
Substituting $\theta_\rho$ from
Eq.~(\ref{EQ:ms_XfourA}) into Eq.~(\ref{EQ:ms_Xfive}) we find
the $T=0$ supercurrent to be
\begin{equation}
j(x)=2v_\rho K_\rho\frac{1}{2\pi}\partial_x \chi.
\label{EQ:ms_Xsix}
\end{equation}
We can confirm this result by considering the case of a Galilean-invariant
system. For such a system we know that
\begin{equation}
j(x)=\rho_s v_s=\rho_s\frac {\hbar}{2m}\partial_x \chi,
\label{EQ:ms_Xseven}
\end{equation}
where $\rho_s$ is the density of superconducting electrons.
At $T=0$ we will have $\rho_s=\rho$.
Comparing Eqs.~(\ref{EQ:ms_Xsix}) and (\ref{EQ:ms_Xseven}),
we see that consistency requires the equilibrium number density in the
Galilean-invariant liquid
to be given by
\begin{equation}
\rho=2K_\rho v_\rho m/\pi\hbar.
\label{EQ:ms_Xeight}
\end{equation}
(The factor of $2$ in this equation arises from the two
spin-projections.)
That this is correct is shown by comparing the commutator
\begin{equation}
[\psi^{\dagger}\psi(x), \frac{\hbar}{2mi}
\psi^{\dagger}(x'){\buildrel\leftrightarrow\over
\partial_{x'}}\psi(x')]=\frac{\hbar}{mi}
\psi^{\dagger}\psi(x)\partial_x \delta(x-x')
\label{EQ:ms_Xnine}
\end{equation}
of the charge and current in a Galilean invariant system with the
corresponding commutator in our Luttinger system, viz.,
\begin{equation}
[\rho(x), j(x')]=-2iK_\rho
v_\rho\frac{1}{\pi}\partial_x\delta(x-x').
\label{EQ:ms_Xten}
\end{equation}
The Luttinger model approximates the Galilean invariant system by the
replacement of the charge density operator on the right hand side of
Eq.~(\ref{EQ:ms_Xnine}) by its expectation value. This confirms that
Eq.~(\ref{EQ:ms_Xeight}) is correct.
We now apply Eq.(\ref{EQ:ms_Xseven}). In the purely LL segment of the line
(i.e., $0<x<L$)
the $\theta_\rho$ and $\phi_\sigma$ fields are no longer constrained by
the condensate. However, as we mentioned earlier, their values at the ends of
the interval are fixed, just as in Eq.~(\ref{eq:bcinf}):
\begin{equation}
\int_{x_1}^{x_2} j(x)dx=2v_\rho
K_\rho\frac{1}{2\pi}(\chi_2-\chi_1)
\label{EQ:ms_Xeleven}
\end{equation}
This is the same result as Eq.~(\ref{eq:jlow}),
because the quantity found by the
thermodynamic trick of differentiating the free energy with respect to
$\chi$ is the spatial average of the current. The advantage of
Eq.~(\ref{EQ:ms_Xeight}) is that we can see that this average current is
independent of the precise way in which the gap goes to zero as we
enter the Luttinger link. Indeed, because the duality map between
one-dimensional charge density waves (CDW) and superconductors interchanges
the charge-
and current-densities, the results we have just described are just
the dual of the well-known result in the theory of
CDW systems that the total charge induced in a region is a
topological quantity depending only on the asymptotic values of the CDW
condensate phase \cite{REF:Gol_Wil}.
\section{Acknowledgements}
We thank Eduardo Fradkin for several useful discussions.
This work was supported by the US NSF under grants DMR89-20538 (DLM) and
DMR94-24511 (MS and PMG), and by the NSERC of Canada (DL).
|
\section{Introduction}
\setcounter{equation}{0}
Several problems of mathematical physics lead to the study of the
scattering
properties of linear ordinary differential equations in a singular limit
\begin{equation}\label{sch}
i\varepsilon \psi'(t)=H(t)\psi(t),\;\;\; t\in\hbox{$\mit I$\kern-.277em$\mit R$}, \;\;\varepsilon\rightarrow 0,
\end{equation}
where the prime denotes the derivative with respect to $t$,
$\psi(t)\in\hbox{$\mit I$\kern-.7em$\mit C$}^n$,
$H(t)\in M_n(\hbox{$\mit I$\kern-.7em$\mit C$})$, for all $t$. A system described by such an equation
will be called a
$n$-level system. Let us mention for example the study of the adiabatic
limit of the time dependent Schr\"odinger equation or the semiclassical
limit of the
one-dimensional multichannel stationary Schr\"odinger equation at
energies above
the potential barriers, to which we will come back below. When the
generator $H(t)$
is well behaved at $+\infty$ and $-\infty$, the scattering properties of
the problem
can be described by means of a matrix naturally associated with equation
(\ref{sch}),
the so-called $S$-matrix. This matrix relates the behavior of
the solution $\psi(t)$
as $t\rightarrow -\infty$ to that of $\psi(t)$ as $t\rightarrow +\infty$. Assuming that
the spectrum
$\sigma(t)$ of $H(t)$ is real and non-degenerate,
\begin{equation}
\sigma(t)=\{e_1(t)<e_2(t)<\cdots <e_n(t)\}\in \hbox{$\mit I$\kern-.277em$\mit R$},
\end{equation}
the $S$-matrix is essentially given by the identity matrix
\begin{equation}
S=\mbox{diag }(s_{11}(\varepsilon),s_{22}(\varepsilon),\cdots s_{nn}(\varepsilon)) +
{\cal O}(\varepsilon^{\infty}),\;\;\;
\mbox{where }s_{jj}(\varepsilon)=1+{\cal O}(\varepsilon)\mbox{ as }\varepsilon\rightarrow 0,
\end{equation}
provided $H(t)$ is $C^{\infty}$, see e.g. \cite{f1}, \cite{f2}, \cite{w}.
Moreover, if $H(t)$ is assumed to
be analytic, it was proven in various situations that the off-diagonal
elements
$s_{jk}$ of $S$ are exponentially decreasing \cite{ff}, \cite{w},
\cite{f1}, \cite{f2}, \cite{jkp}, \cite{jp3}
\begin{equation}
s_{jk}={\cal O} \left(\mbox{e}^{-\kappa /\varepsilon}\right),\;\;\; \forall j\neq k,
\end{equation}
as $\varepsilon\rightarrow 0$.
See also \cite{jp4}, \cite{n}, \cite{m}, \cite{sj} for corresponding
results in infinite dimensional spaces.
Since the physical information is often contained in these off-diagonal
elements, it is of interest to be able to give an asymptotic formula for
$s_{jk}$, rather
then a mere estimate.
For $2$-level systems (or systems reducible to this case, see \cite{jp5},
\cite{j1}, \cite{mn}), the situation is now
reasonably well understood, at least under generic circumstances. Indeed,
a rigorous study of the $S$-matrix associated with (\ref{sch}) when $n=2$,
under the hypotheses
loosely stated above, is provided in the recent paper \cite{jp3}. The
treatment presented unifies
in particular earlier results obtained either for the time dependent
adiabatic schr\"odinger equation, see e.g \cite{jp7} and references
therein, or for
the study of the above barrier reflexion in the semiclassical limit, see
e.g. \cite{ff},\cite{o}.
Further references are provided in \cite{jp3}. As an intermediate result,
the asymptotic formula
\begin{equation}\label{exaf}
s_{jk}=g_{jk}\mbox{e}^{-\Gamma_{jk}/\varepsilon}\left(1+{\cal O} (\varepsilon)\right),\;\;\;\varepsilon
\rightarrow 0,
\end{equation}
for $j\neq k\in\{1,2\}$, with $g_{jk}\in\hbox{$\mit I$\kern-.7em$\mit C$}$ and $\mbox{Re }\Gamma_{jk}>0$
is proven in \cite{jp3}.
As is well known, to get an asymptotic formula for $s_{jk}$, one has to
consider (\ref{sch}) in the complex plane, in particular in the vicinity
of the degeneracy points of the
analytic continuations of eigenvalues
$\mbox{e}_1(z)$ and $e_2(z)$. {\em Provided} the level lines of the multivalued
function
\begin{equation}\label{lln}
\mbox{Im }\int_0^ze_1(z')-e_2(z')dz'= \mbox{cst},
\end{equation}
called Stokes lines, naturally associated with (\ref{sch}) behave
properly in the
complex plane, the so-called complex WKB method allows to prove
(\ref{exaf}).
But more important, it is also shown in \cite{jp3} how to improve
(\ref{exaf})
to an asymptotic formula accurate up to exponentially small relative
error
\begin{equation}\label{exasup}
s_{jk}=g_{jk}^*(\varepsilon)\mbox{e}^{-\Gamma_{jk}^*(\varepsilon)/\varepsilon}\left(1+{\cal O}
\left(\mbox{e}^{-\kappa /\varepsilon}\right)\right),
\;\;\;\varepsilon\rightarrow 0,
\end{equation}
with $g_{jk}^*(\varepsilon)=g_{jk}+{\cal O} (\varepsilon)$ and $\Gamma_{jk}^*(\varepsilon)=
\Gamma_{jk}+{\cal O}(\varepsilon^2)$.
This is achieved by
using a complex WKB analysis jointly with the recently developed
superasymptotic theory
\cite{be}, \cite{n}, \cite{jp5}.
Note that when given a generator,
the principal difficulty in justifying formulas (\ref{exaf}) and
(\ref{exasup}) is the verification
that the corresponding Stokes
lines (\ref{lln}) display the proper behavior {\em globally} in the
complex plane, which may or may not be the case
\cite{jkp}. However, this condition is always satisfied when the complex
eigenvalue degeneracy is close
to the real axis, as shown in \cite{j1}. See also \cite{mn} and \cite{r}
for recent related results.
For $n$-level systems, with $n\geq 3$, the situation is by no means as
well understood.
There are some results obtained with particular generators. In \cite{de},
\cite{ch1},
\cite{ch2} and \cite{bve}, certain elements of the $S$-matrix are
computed if $H(t)=H^*(t)$ depends
linearly on $t$, $H(t)=A+tB$,
for some particular matrices $A$ and $B$. The choices of $A$ and $B$ are
such that all
components of the solution
$\psi(t)$ can be deduced from the first one and an exact integral
representation of this
first component can be obtained. The integral representation is analyzed
by standard asymptotic
techniques and this leads to results which are valid for any $\varepsilon>0$, as
for the classical
Landau-Zener generator.
The study of the three-level problem when $H(t)=H^*(t)\in M_3(\hbox{$\mit I$\kern-.277em$\mit R$})$ is
tackled in the closing
section of the very interesting paper \cite{hp}.
A non rigorous and essentially local discussion
of the behavior of the level lines $\mbox{Im }\int_0^ze_j(z')-e_k(z')dz'$,
$j\neq k=1,2,3$, is provided
and justifies in very favorable cases an asymptotic formula for some
elements of the
$S$-matrix. See also the review \cite{s}, where a non rigorous study of
(\ref{sch})
is made close to a complex degeneracy point of a group of eigenvalues by
means of an exact solution
to a model equation.
However, no asymptotic formula for $s_{jk}$, $j\neq k$, can be
found in the literature for general $n$-level systems, $n\geq 3$. This is
due to the fact
that the direct generalization of the method used successfully for
$2$-level
systems may lead to seemingly inextricable difficulties for $n=3$
already. Indeed, with three eigenvalues,
one has to consider three sets of level lines $\mbox{Im }
\int_0^ze_j(z')-e_k(z')dz'$ to
deal with (\ref{sch}) in the complex plane, and the conditions they have
to fulfill in order that
the limit $\varepsilon\rightarrow 0$ can be controlled may
be incompatible for a given generator, see \cite{f1}, \cite{f2} and
\cite{hp}. It should be mentioned however, that
there are specific examples in which this difficult problem can be
mastered. Such a result
was recently obtained in the semiclassical study \cite{ba} of a particular
problem of
resonances for which similar considerations in the complex plane are
required.
The goal of this paper is to provide some general insight on the
asymptotic computation of the
$S$-matrix associated with $n$-level systems, $n\geq 3$, based on a
generalization of the techniques which proved
to be successful for $2$-level systems.
The content of this paper is twofold. On the one hand we set up a general
framework in which
asymptotic formulas for the exponentially small off-diagonal coefficients
can be proven. On the other hand we actually prove such formulas for a
wide class of $n$-level systems.
In the first part of the paper, we give our definition of the $S$-matrix
associated with
equation (\ref{sch}) and explicit the symmetries it inherits from the
symmetries of $H(t)$,
for $t\in\hbox{$\mit I$\kern-.277em$\mit R$}$ (proposition \ref{PROS}). Then we
turn to the determination of the analyticity properties of the
eigenvalues
and eigenvectors of $H(z)$, $z\in\hbox{$\mit I$\kern-.7em$\mit C$}$, which are at the root of the
asymptotic formulas we derive
later (lemma \ref{ECHCO}). The next step is the formulation of sufficient
conditions adapted to the scattering situation
we consider, under which a complex WKB analysis allows to prove a formula
like (\ref{exaf})
(proposition \ref{eslf}). The
conditions stated are similar but not identical to those given in
\cite{jkp} or \cite{hp}.
As a final step, we show how to improve the asymptotic formula
(\ref{exaf}) to (\ref{exasup}) by means of the superasymptotic machinery
(proposition \ref{MAPROP} and
lemma \ref{sdec}).
Then we turn to the second part of the paper, where we show that a wide
class of
generators fits into our framework and satisfies our conditions.
These generators are obtained by perturbation of
generators whose eigenvalues display degeneracies on the real axis, in
the spirit of \cite{j1}.
We prove that for these generators, in absence of any symmetry of the
generator $H(t)$, one element
per column at least in the $S$-matrix can be
asymptotically computed (theorem \ref{PERCO}).
This is the main technical section of the paper. The major advantage of
this
construction is that it is sufficient to look at the behavior of the
eigenvalues on the real axis to
check if the conditions are satisfied. The closing section contains an
application of our general
results to the study of quantum adiabatic transitions in the time
dependent Schr\"odinger equation
and of the semiclassical scattering properties of the multichannel
stationary Schr\"odinger equation. In particular, explicit use of the
symmetries of the $S$-matrix is
made to increase the number of its elements for which an asymptotic
formula holds.
In the latter application, further specific symmetry properties
of the $S$-matrix are derived (lemma \ref{REPOT}).
\vspace{.5cm} \\
{\bf Acknowledgments:} It is a great pleasure to thank Charles-Edouard
Pfister
for many enlightening and fruitful discussions which took place in
Marseille and
Lausanne. The hospitality of the Institut de Physique Th\'eorique de
l'EPFL
where part of this work was accomplished is acknowledged.
\section{Definition and properties of the $S$-matrix.}
\setcounter{equation}{0}
We consider the evolution equation
\begin{equation}\label{schr}
i\varepsilon \psi'(t)=H(t)\psi(t),\;\;\; t\in\hbox{$\mit I$\kern-.277em$\mit R$}, \;\;\varepsilon\rightarrow 0,
\end{equation}
where the prime denotes the derivative with respect to $t$,
$\psi(t)\in\hbox{$\mit I$\kern-.7em$\mit C$}^n$,
$H(t)\in M_n(\hbox{$\mit I$\kern-.7em$\mit C$})$, for all $t$.
We make some assumptions on the
generator $H(t)$. The first one is the usual analyticity condition in this
context.\\
{\bf H1}{\em
There exists a strip
\begin{equation}
S_{\alpha}=\left\{z\in\hbox{$\mit I$\kern-.7em$\mit C$} |\, |\mbox{\em Im }z|\leq\alpha\right\},\;\;
\alpha>0,
\end{equation}
such that $H(z)$ is analytic for all $z\in S_{\alpha}$.}\\ Since we are
studying scattering properties, we need sufficient decay at infinity.\\
{\bf H2} {\em There exist two nondegenerate matrices $H(+), H(-) \in
M_n(\hbox{$\mit I$\kern-.7em$\mit C$})$
and $a>0$ such that
\begin{equation}
\lim_{t\rightarrow\pm \infty}|t|^{1+a}\sup_{|s|\leq\alpha}\|H(t+is)-H(\pm)\|
<\infty.
\end{equation}}
We finally give a condition which has to do with the physics behind
the problem.\\
{\bf H3} {\em
For $t\in\hbox{$\mit I$\kern-.277em$\mit R$}$, the spectrum of $H(t)$, denoted by $\sigma(t)$,
is real and non-degenerate
\begin{equation}
\sigma(t)=\left\{e_1(t)<e_2(t)<\cdots <e_n(t)\right\}\subset \hbox{$\mit I$\kern-.277em$\mit R$}
\end{equation}
and there exists $g>0$ such that
\begin{equation}
\inf_{j\neq k \atop t\in \hbox{$\mit I$\kern-.277em$\mit R$}}\left|e_j(t)-e_k(t)\right|\geq g.
\end{equation}
}
As a consequence of H3, for each $t\in \hbox{$\mit I$\kern-.277em$\mit R$}$, there exists a complete set
of projectors $P_j(t)=P_j^2(t)\in M_n(\hbox{$\mit I$\kern-.7em$\mit C$})$, $j=1,2,\cdots ,n$ such that
\begin{eqnarray}
& &\sum_{j=1}^n P_j(t)\equiv \hbox{$\mit I$\kern-.72em$\mit I$}\\
& &H(t)=\sum_{j=1}^n e_j(t)P_j(t)
\end{eqnarray}
and there exists a basis of $\hbox{$\mit I$\kern-.7em$\mit C$}^n$ of eigenvectors of
$H(t)$. We determine these eigenvectors $\varphi_j(t)$, $j=1,2,\cdots ,n$
uniquely (up to a constant) by requiring them to satisfy
\begin{eqnarray}
& &H(t)\varphi_j(t)=e_j(t)\varphi_j(t)\\
& &P_j(t)\varphi_j'(t)\equiv 0, \;\;\;\;\;\;\;\;\;\;j=1,2,\cdots ,n.
\label{phco}
\end{eqnarray}
Explicitly, if $\chi_j(t)$, $j=1,2,\cdots ,n$ form a complete set of
differentiable eigenvectors of $H(t)$, the eigenvectors
\begin{equation}
\varphi_j(t)=\mbox{e}^{-\int_0^t\xi_j(t')dt'}\psi_j(t),\;\;\;\mbox{s.t. }\varphi_j(0)
=\psi_j(0)
\end{equation}
with
\begin{equation}
\xi_j(t)=\frac{\langle\psi_j(t)|P_j(t)\psi_j'(t)\rangle}{\|\psi_j(t)\|^2},
\;\;\;
j=1,\cdots ,n
\end{equation}
verify (\ref{phco}).
That this choice leads to an analytic set of eigenvectors close to the
real axis will be proven below.
We expand the solution $\psi(t)$ along the basis just constructed, thus
defining unknown coefficients $c_j(t)$, $j=1,2,\cdots ,n$ to be determined,
\begin{equation}\label{expa}
\psi(t)=\sum_{j=1}^nc_j(t)\mbox{e}^{-i\int_0^te_j(t')dt'/\varepsilon }\varphi_j(t).
\end{equation}
The phases $\mbox{e}^{-i\int_0^te_j(t')dt'/\varepsilon }$ (see H3) are introduced for
convenience. By inserting (\ref{expa}) in (\ref{schr}) we get the
following differential equation for the $c_j(t)$'s
\begin{equation}\label{eqco}
c_j'(t)=\sum_{k=1}^na_{jk}(t)\mbox{e}^{i\Delta_{jk}(t)/\varepsilon}c_k(t)
\end{equation}
where
\begin{equation}
\Delta_{jk}(t)=\int_0^t(e_j(t')-e_k(t'))dt'
\end{equation}
and
\begin{equation}\label{coup}
a_{jk}(t)=-\frac{\langle\varphi_j(t)|P_j(t)\varphi_k'(t)\rangle}{\|\varphi_j(t)\|^2}.
\end{equation}
Here $\langle \cdot |\cdot \rangle$ denotes the usual scalar product in $\hbox{$\mit I$\kern-.7em$\mit C$}^n$.
Our choice (\ref{phco}) implies $a_{jj}(t)\equiv 0$. It is also
shown below that the $a_{jk}(t)$'s are analytic functions in a
neighborhood
of the real axis and that hypothesis H2 imply that they satisfy the
estimate
\begin{equation}
\lim_{t\rightarrow\pm\infty}\sup_{j\neq k}
|t|^{1+a}\left|a_{jk}(t)\right|<\infty.
\end{equation}
As a consequence of this last property and of the fact that the
eigenvalues
are real by assumption, the following limits exist
\begin{equation}
\lim_{t\pm\infty}c_j(t)=c_j(\pm\infty).
\end{equation}
We are now able to define the associated $S$-matrix, $S\in M_n(\hbox{$\mit I$\kern-.7em$\mit C$})$, by
the
identity
\begin{equation}
S\pmatrix{c_1(-\infty)\cr c_2(-\infty) \cr \vdots \cr c_n(-\infty)}=
\pmatrix{c_1(+\infty)\cr c_2(+\infty) \cr \vdots \cr c_n(+\infty)}.
\end{equation}
Such a relation makes sense because of the linearity of the equation
(\ref{eqco}).
It is a well known result that under our general hypotheses, the
$S$-matrix
satisfies
\begin{equation}
S=\hbox{$\mit I$\kern-.72em$\mit I$} +{\cal O} (\varepsilon).
\end{equation}
Note that the $j^{\mbox{\scriptsize th}}$ column of the $S$-matrix
is given by the solution of (\ref{eqco}) at $t=\infty$
subjected to the initial
conditions $c_k(-\infty)=\delta_{jk}$, $k=1,2,\cdots ,n$.
In general, the $S$-matrix defined above has no particular properties
besides
that of being invertible. However, when the generator $H(t)$ satisfies
some
symmetry properties, the same is true for $S$. As such properties are
important in applications, we show below that if $H(t)$ is self-adjoint
with respect to some indefinite scalar product, then
$S$ is unitary with respect to another indefinite scalar product.
Let $J\in M_n(\hbox{$\mit I$\kern-.7em$\mit C$})$ be an invertible self-adjoint matrix.
We define an indefinite metric on $\hbox{$\mit I$\kern-.7em$\mit C$}^n$ by means of the indefinite scalar
product
\begin{equation}
(\cdot ,\cdot )_J=\langle \cdot |J\cdot \rangle.
\end{equation}
It is easy to check that the adjoint $A^{\#}$ of a matrix $A$ with
respect to the
$(\cdot ,\cdot )_J$ scalar product is given by
\begin{equation}
A^{\#}=J^{-1}A^*J.
\end{equation}
\begin{prop}\label{PROS}
Let $H(t)$ satisfy H1, H2 and possess $n$ distinct eigenvalues $\forall
t\in\hbox{$\mit I$\kern-.277em$\mit R$}$.
Further assume $H(t)$ is self-adjoint with respect to the
scalar product $(\cdot ,\cdot )_J$,
\begin{equation}
H(t)=H^{\#}(t)=J^{-1}H^*(t)J,\;\;\; \forall t\in \hbox{$\mit I$\kern-.277em$\mit R$},
\end{equation}
and the eigenvectors $\varphi_j(0)$ of $H(0)$ satisfy
\begin{equation}
(\varphi_j(0) ,\varphi_j(0) )_J=\rho_j,\;\;\;\rho_j\in\{-1,1\},\;\;\;\forall
j=1,\cdots, n.
\end{equation}
Then the eigenvalues of $H(t)$ are real $\forall t\in\hbox{$\mit I$\kern-.277em$\mit R$}$ and the
$S$-matrix is
unitary with respect to the scalar product
$(\cdot ,\cdot )_R$, where $R=R^*=R^{-1}$ is the real diagonal matrix
$R=\mbox{diag }(\rho_1,\rho_2,\cdots ,\rho_n)$,
\begin{equation}
S^{\#}=RS^*R=S^{-1}.
\end{equation}
\end{prop}
{\bf Remark:}\\
The condition $(\varphi_j(0) ,\varphi_j(0) )_J=\pm 1$ can always be satisfied by
suitable renormalization provided $(\varphi_j(0) ,\varphi_j(0) )_J\neq 0$.
The main interest of this proposition is that when the $S$-matrix
possesses
symmetries, some of its elements can be deduced from
resulting identities, without resorting to their actual computations.
A simple proof of the proposition making use of notions discussed in the
next section
can be found in appendix.
The above proposition can
actually be used for the two main applications we deal with in section
\ref{applic}.
Note that in specific cases, further symmetry property can be derived for
the $S$-matrix,
see section \ref{applic}.
\section{Analyticity properties}\label{anprop}
\setcounter{equation}{0}
The generator $H(z)$ is analytic in $S_{\alpha}$, hence the solution of
the
linear equation (\ref{schr}) $\psi(z)$ is analytic in $S_{\alpha}$ as
well.
However, the eigenvalues and eigenprojectors of $H(z)$ may have
singularities in $S_{\alpha}$. Let us recall some basic
properties, the proofs of which can be found in \cite{k}. The eigenvalues
and eigenprojectors of a matrix analytic in a region of the complex plane
have analytic continuations in that region with
possible singularities located at points $z_0$ called exceptional points.
In a neighborhood free of exceptional points,
the eigenvalues are given by branches of analytic functions and their
multiplicities are constant.
One eigenvalue can therefore be analytically continued until it
coincides at
$z_0$ with one or several other eigenvalues. The set of such points
defines the set of exceptional points.
The eigenvalues may possess branching points at an
exceptional $z_0$, where they are continuous, whereas the eigenprojectors
are
also multivalued but diverge as $z\rightarrow z_0$. Hence, by hypothesis H3,
the $n$ distinct eigenvalues $e_j(t)$
defined on the real axis are analytic on the real axis and possess
multivalued analytic
continuations in $S_{\alpha}$, with possible branching points at the set
of
degeneracies $\Omega$, given by
\begin{equation}
\Omega=\left\{ z_0 |\; e_j(z_0)=e_k(z_0), \mbox{for some $k,j$ and
some
analytic continuation} \right\}.
\end{equation}
By assumption H2, $\Omega$ is finite, by H3, $\Omega\cap\hbox{$\mit I$\kern-.277em$\mit R$}=\emptyset$ and
$\Omega=\overline{\Omega}$, due to Schwarz's principle. Similarly, the
eigenprojectors $P_j(t)$ defined on the real axis are analytic on the real
axis and possess multivalued analytic continuations in $S_{\alpha}$, with
possible singularities at $\Omega$.
To see more precisely what happens to these multivalued functions
when we turn around a point $z_0\in\Omega$, we consider the
construction described in figure \ref{fig1}. Let $f$ be a multivalued
analytic function in
$S_{\alpha}\backslash\Omega$. We denote by $f(z)$ the analytic
continuation
of $f(0)$ along some path $\beta\in S_{\alpha}\backslash\Omega$ from $0$
to $z$.
Then we perform the analytic continuation of $f(z)$ along a negatively
oriented loop $\delta$ based at $z$ around a unique point $z_0\in\Omega$,
and denote by $\widetilde{f}(z)$ the function we get when we come back at
the starting point (if $\delta$ is positively oriented,
the construction is similar). For later purposes, we define $\eta_0$ as
the
negatively
oriented loop homotopic to the loop based at the origin encircling $z_0$
obtained by following
$\beta$ from $0$ to $z$, $\delta$ from $z$ back to $z$ and $\beta$ in the
reverse sense from $z$ back to the origin.
\begin{figure}
\vspace{5cm}
\hspace{3cm}\special{picture fig1 scaled 700}
\caption{The paths $\beta$, $\delta$ and $\eta_0$ in
$S_{\alpha}\backslash \Omega$.}\label{fig1}
\end{figure}
We'll keep this notation in the rest of this section.
It follows from the foregoing that if we perform the analytic continuation
of the set of eigenvalues $\{e_j(z)\}_{j=1}^n$, along a negatively
oriented loop around $z_0\in\Omega$, we get the set
$\{\widetilde{e}_j(z)\}_{j=1}^n$ with
\begin{equation}
\widetilde{e}_j(z)=e_{\sigma_0(j)}(z), \;\;j=1,\cdots,n,
\end{equation}
where
\begin{equation}
\sigma_0 :\;\{1,2,\cdots,n\}\rightarrow \{1,2,\cdots,n\}
\end{equation}
is a permutation which depends on $\eta_0$. Similarly, and with the same
notations, we get for the analytic continuations of the projectors around
$z_0$
\begin{equation}\label{prop}
\widetilde{P}_j(z)=P_{\sigma_0(j)}(z), \;\;j=1,\cdots,n.
\end{equation}
Let us consider now the eigenvectors $\varphi_j(t)$. We define $W(t)$ as the
solution of
\begin{eqnarray}\label{par}
W'(t)&=&\sum_{j=1}^nP_j'(t)P_j(t)W(t)\\
&\equiv&K(t)W(t),\;\;\; W(0)=\hbox{$\mit I$\kern-.72em$\mit I$},\nonumber
\end{eqnarray}
where $t\in\hbox{$\mit I$\kern-.277em$\mit R$}$. It is well known \cite{k}, \cite{kr}, that $W(t)$
satisfies the
intertwining identity
\begin{equation}\label{inter}
W(t)P_j(0)=P_j(t)W(t),\;\;\;j=1,2,\cdots,n\; ,\;\forall t\in\hbox{$\mit I$\kern-.277em$\mit R$},
\end{equation}
so that, if
$\{\varphi_j(0)\}_{j=1}^n$
denotes a set of eigenvectors of $H(0)$, the vectors defined by
\begin{equation}\label{defeig}
\varphi_j(t)=W(t)\varphi_j(0)
\end{equation}
are eigenvectors of $H(t)$. Moreover, using the identity
$Q(t)Q'(t)Q(t)\equiv 0$ which is true for any differentiable projector,
it
is easily checked that condition (\ref{phco}) is satisfied by these
vectors. The generator $K(t)$ is analytic on the real axis and can be
analytically continued in $S_{\alpha}\backslash \Omega$. Actually, $K(z)$
is single valued in $S_{\alpha}\backslash \Omega$. Indeed, let us consider
the analytic continuation of $K(z)$ around
$z_0\in\Omega$. We get from (\ref{prop}) that
\begin{equation}
\widetilde{P}_j'(z)=P_{\sigma_0(j)}'(z),
\end{equation}
so that
\begin{eqnarray}
\widetilde{K}(z)&=&\sum_{j=1}^n\widetilde{P}_j'(z)\widetilde{P}_j(z)
=\sum_{j=1}^n{P}_{\sigma_0(j)}'(z){P}_{\sigma_0(j)}(z)
\nonumber\\
&=&\sum_{k=1}^n{P}_{k}'(z){P}_{k}(z)=K(z).
\end{eqnarray}
Consequently, $W(t)$ can be analytically continued in
$S_{\alpha}\backslash \Omega$, where it is multivalued and satisfies
both
(\ref{par}) and (\ref{inter}) with $z\in S_{\alpha}\backslash \Omega$ in
place
of $t\in\hbox{$\mit I$\kern-.277em$\mit R$}$. Moreover, the relation between the analytic continuation
$W(z)$
from $0$ to some point $z\in S_{\alpha}\backslash \Omega$
and the analytic continuation $\widetilde{W}(z)$ is given by a monodromy
matrix $W(\eta_0)$ such that
\begin{equation}\label{mono}
\widetilde{W}(z)=W(z)W(\eta_0),
\end{equation}
where $\eta_0$ is the negatively oriented loop based at the origin
which encircles $z_0\in\Omega$ only, (see figure
\ref{fig1}).
Note also that the analytic continuation $W(z)$ is invertible in
$S_{\alpha}\backslash \Omega$ and $W^{-1}(z)$ satisfies
\begin{equation}
{W^{-1}}'(z)=-W^{-1}(z)K(z),\;\;\; W^{-1}(0)=\hbox{$\mit I$\kern-.72em$\mit I$}.
\end{equation}
As a consequence, the eigenvectors (\ref{defeig}) possess multivalued
analytic extensions
in $S_{\alpha}\backslash \Omega$. Consider the relation
\begin{equation}
H(z)\varphi_j(z)=e_j(z)\varphi_j(z), \;\;
\end{equation}
obtained by analytic continuation from $0$ to some point
$z\in S_{\alpha}\backslash \Omega$. When analytically
continued along a negatively oriented loop around $z_0\in\Omega$, it
yields
\begin{equation}
\widetilde{H}(z)\widetilde{\varphi}_j(z)=H(z)\widetilde{\varphi}_j(z)=
\widetilde{e}_j(z)\widetilde{\varphi}_j(z)=e_{\sigma_0(j)}(z)
\widetilde{\varphi}_j(z).
\end{equation}
Thus $\widetilde{\varphi}_j(z)$ is proportional to ${\varphi}_{\sigma_0(j)}(z)$
and
we introduce the quantity $\theta_j(\eta_0)\in\hbox{$\mit I$\kern-.7em$\mit C$}$ by the definition
\begin{equation}\label{theta}
\widetilde{\varphi}_j(z)=\mbox{e}^{-i\theta_j(\eta_0)}{\varphi}_{\sigma_0(j)}(z),
\;\;\;j=1,2,\cdots,n\;.
\end{equation}
This is equivalent to (see (\ref{mono}))
\begin{equation}
W(\eta_0)\widetilde{\varphi}_j(0)=\mbox{e}^{-i\theta_j(\eta_0)}
{\varphi}_{\sigma_0(j)}(0).
\end{equation}
Let us consider the couplings (\ref{coup}). Using the definition
(\ref{defeig}),
the invertibility of $W(t)$ and the identity (\ref{inter}), it's not
difficult
to see that we can rewrite
\begin{equation}\label{anajk}
a_{jk}(t)=-\frac{\langle\varphi_j(0)|P_j(0)W(t)^{-1}K(t)W(t)\varphi_k(0)\rangle}
{\|\varphi_j(0)\|^2},\;\;\; t\in\hbox{$\mit I$\kern-.277em$\mit R$},
\end{equation}
which is analytic on the real axis and can be analytically continued in
$S_{\alpha}\backslash \Omega$, where it is multivalued. Thus, the same
is true
for the coefficients $c_j(t)$ which satisfy the linear differential
equation
(\ref{eqco}) and their analytic continuations satisfy the same equation
with
$z\in S_{\alpha}\backslash \Omega$ in place of $t\in \hbox{$\mit I$\kern-.277em$\mit R$}$. We now come to
the
main identity of this section, regarding the coefficients $c_j(z)$. Let
us denote
by $c_j(z)$ the analytic continuation of $c_j(0)$ from $0$ to some
$z\in S_{\alpha}\backslash \Omega$. We perform
the analytic continuation of $c_j(z)$ along a negatively oriented loop
around
$z_0\in\Omega$ and denote by $\widetilde{c}_j(z)$ the function we get
when
we come back at the starting point $z$.
\begin{lem}\label{ECHCO}
For any $j=1,\cdots,n$, we have
\begin{equation}\label{echco}
\widetilde{c}_j(z)\mbox{e}^{-i\int_{\eta_0}e_j(u)du/\varepsilon}
\mbox{e}^{-i\theta_j(\eta_0)}
=c_{\sigma_0(j)}(z)
\end{equation}
where $\eta_0$, $\theta_j(\eta_0)$ and $\sigma_0(j)$ are defined as above.
\end{lem}
{\bf Proof:}\\
It follows from hypothesis H1 that $\psi(z)$ is analytic in $S_{\alpha}$
so that
\begin{eqnarray}
& &\sum_{j=1}^nc_j(z)\mbox{e}^{-i\int_0^ze_j(u)du/\varepsilon }\varphi_j(z)=
\sum_{j=1}^n\widetilde{c}_j(z)\widetilde{\mbox{e}^{-i\int_0^ze_j(u)du/\varepsilon }}
\widetilde{\varphi}_j(z)=\nonumber\\
& &\sum_{j=1}^n\widetilde{c}_j(z){\mbox{e}^{-i\int_{\eta_0}e_j(u)du/\varepsilon }}
\mbox{e}^{-i\int_0^ze_{\sigma_0(j)}(u)du/\varepsilon}\mbox{e}^{-i\theta_j(\eta_0)}
{\varphi}_{\sigma_0(j)}(z).
\end{eqnarray}
We conclude by the fact that $\{\varphi_j(z)\}_{j=1}^n$ is a basis.\hfill {$\Box$} \\
{\bf Remark:} \\
It is straightforward to generalize the study of the analytic
continuations around one singular point of the functions given above
to the case where the analytic continuations are performed around several
singular points, since $\Omega$ is finite. The loop $\eta_0$ can be
rewritten
as a finite succession of individual loops encircling one point of
$\Omega$
only, so that the permutation $\sigma_0$ is given by the composition of
a finite number of individual permutations. Thus
the factors $\mbox{e}^{-i\theta_j(\eta_0)}$ in (\ref{theta}) should be replaced
by a
product of such factors, each associated with one individual loop and the
same
is true for the factors $\exp(-i\int_{\eta_0}e_j(z)dz/\varepsilon )$ in lemma
\ref{ECHCO}. This process is performed in the proof of theorem
\ref{PERCO}.
\section{Complex WKB analysis}
\setcounter{equation}{0}
This section is devoted to basic estimates on the coefficients $c_j(z)$
in
certain domains extending to infinity in both the positive and negative
directions
inside the strip $S_{\alpha}$.
We first consider what happens in
neighborhoods of $\pm\infty$.
It follows from assumption H1 by a direct application of the Cauchy
formula
that
\begin{equation}
\lim_{t\rightarrow\pm\infty}\sup_{|s|\leq\alpha}|t|^{1+a}\|H'(t+is)\|<\infty.
\end{equation}
Hence the same is true for the single valued matrix $K(z)$
\begin{equation}\label{deck}
\lim_{t\rightarrow\pm\infty}\sup_{|s|\leq\alpha}|t|^{1+a}\|K(t+is)\|<\infty.
\end{equation}
Let $0<T\in\hbox{$\mit I$\kern-.277em$\mit R$}$ be such that
\begin{equation}\label{deft}
\min_{z\in\Omega}\mbox{Re }z>-T,\;\;\;\mbox{and }
\max_{z\in\Omega}\mbox{Re }z<+T.
\end{equation}
All quantities encountered so far are analytic in
$S_{\alpha}\cap\{z||\mbox{Re }z|>T\}$, and we denote by a
"$\widetilde{\hspace{.2cm}}$" any
analytic continuation in that set. As noticed earlier
\begin{equation}
\widetilde{W}'(z)=K(z)\widetilde{W}(z),\;\;\; z\in S_{\alpha}
\cap\{z||\mbox{Re }z|>T\}
\end{equation}
so that it follows from (\ref{deck}) that the limits
\begin{equation}
\lim_{t\rightarrow\pm\infty}\widetilde{W}(t+is)=\widetilde{W}(\pm\infty)
\end{equation}
exist uniformly in $s\in ]-\alpha,\alpha[$.
Consequently, see (\ref{anajk}),
\begin{equation}\label{refaj}
\lim_{t\rightarrow\pm\infty}|t|^{1+a}\sup_{|s|\leq\alpha}|\widetilde{a}_{jk}
(t+is)|<\infty,
\;\;\;\forall j,k\in\{1,\cdots,n\}.
\end{equation}
Finally, for $|t|>T$, we can write
\begin{eqnarray}
\mbox{Im }\widetilde{\Delta}_{jk}(t+is)&=&\mbox{Im }\left(
\int_{\eta}e_j(z)dz
-\int_{\eta}e_k(z)dz\right)\nonumber\\
&+&\int_0^s\mbox{Re }(e_{\sigma_j(j)}(t+is')-
e_{\sigma_j(k)}(t+is'))ds',
\end{eqnarray}
where this equation is obtained by deforming the path of integration
from $0$ to
$z=t+is$ into a loop $\eta$ based at the origin, which may encircle
points of
$\Omega$, followed by the
real axis from $0$ to $\mbox{Re }z$ and a vertical path from
$\mbox{Re }z$
to $z$, see figure \ref{figlo}.
\begin{figure}
\vspace{4cm}
\hspace{3cm}\special{picture fig2 scaled 700}
\caption{The path of integration for $\widetilde{\Delta}_{jk}(z)$
(the x's denote points of $\Omega$).}\label{figlo}
\end{figure}
Hence we have
\begin{equation}
\sup_{z\in S_{\alpha}\cap\{z||\mbox{\scriptsize Re }z|>T\}}
\mbox{Im }\widetilde{\Delta}_{jk}(z)<\infty,
\end{equation}
which, together with (\ref{refaj})
yields the existence of the limits
\begin{equation}\label{unili}
\lim_{t\rightarrow\pm\infty}\widetilde{c}_j(t+is)=\widetilde{c}_j(\pm\infty)
\end{equation}
uniformly in $s\in ]-\alpha,\alpha[$.
We now define the domains in which useful estimates can be obtained.
\\
{\bf Definition:} {\em
Let $j\in\{1,\cdots,n\}$ be fixed.
A {\em dissipative domain for the index $j$},
$D_j\subset S_{\alpha}\backslash \Omega$, is such that
\begin{equation}
\sup_{z\in D_j}\mbox{\em Re }z=\infty,\;\;\;\inf_{z\in D_j}
\mbox{\em Re }z=-\infty,
\end{equation}
and is defined by the property that for any $z\in D_j$ and any
$k\in \{1,\cdots,n\}$, there exists a path
$\gamma^k\subset D_j$ parameterized by $u\in ]-\infty,t]$
which links $-\infty$ to $z$
\begin{equation}
\lim_{u\rightarrow -\infty}\mbox{\em Re }\gamma^k(u)=-\infty,\;\;\;\gamma^k(t)
=z
\end{equation}
with
\begin{equation}
\sup_{z\in D_j}\sup_{u\in ]-\infty,t]}\left|\frac{d}{du}
\gamma^k(u)\right|<\infty
\end{equation}
and satisfies the monotonicity condition
\begin{equation}
\mbox{\em Im }\widetilde{\Delta}_{jk}(\gamma^k(u))\;\;\mbox{is a non
decreasing
function of }u\in]-\infty, t].
\end{equation}
Such a path is a {\em dissipative path for $\{jk\}$}.
Here $\widetilde{\Delta}_{jk}(z)$ is the analytic continuation of
\begin{equation}
{\Delta}_{jk}(t)=\int_0^t(e_j(t')-e_k(t'))dt',\;\;\; t\in\hbox{$\mit I$\kern-.277em$\mit R$},
\end{equation}
in $D_j$ along a path $\beta$ described in figure \ref{fig2} going from
$0$ to
$-T\in\hbox{$\mit I$\kern-.277em$\mit R$}$ along the real axis and
then vertically up or down until it reaches $D_j$, where $T>0$ is chosen
as in
(\ref{deft}).
}\\
{\bf Remark:}\\
The finiteness of $\Omega$ insures the existence of such a path $\beta$.
\begin{figure}
\vspace{5cm}
\hspace{3cm}\special{picture fig3 scaled 700}
\caption{The path $\beta$ along which the analytic continuation of
${\Delta}_{jk}(t)$
in $D_j$ is taken.}\label{fig2}
\end{figure}
Let $\widetilde{c}_k(z)$, $k=1,2,\cdots,n$, $z\in D_j$, be the analytic
continuations
of $c_k(t)$ along the same path $\beta$ which are solutions of the
analytic continuation of
(\ref{eqco}) in $D_j$ along $\beta$
\begin{equation}\label{aneqc}
\widetilde{c}_k'(z)=\sum_{l=1}^n\widetilde{a}_{kl}(z)
\mbox{e}^{i\widetilde{\Delta}_{kl}(z)
/\varepsilon}\widetilde{c}_l(z).
\end{equation}
We take as initial conditions in $D_j$
\begin{equation}\label{coin}
\lim_{\mbox{\scriptsize Re }z\rightarrow -\infty}\widetilde{c}_k(z)=
\lim_{t\rightarrow -\infty}c_k(t)=\delta_{jk},
\;\;\;k=1,\cdots,n.
\end{equation}
and we define
\begin{equation}\label{defx}
x_k(z)=\widetilde{c}_k(z)\mbox{e}^{i\widetilde{\Delta}_{jk}(z)/\varepsilon},\;\;\;
z\in D_j,
\;k=1,\cdots,n.
\end{equation}
\begin{lem}\label{wkb}
In a dissipative domain for the index $j$ we get the estimates
\begin{eqnarray}
& &\sup_{z\in D_j}|x_j(z)-1|={\cal O} (\varepsilon)\\
& &\sup_{z\in D_j}|x_k(z)|={\cal O} (\varepsilon),\;\;\;\forall k\neq j.
\end{eqnarray}
\end{lem}
{\bf Remark:}\\
The real axis is a dissipative domain for all indices. In this case we
have
$\widetilde{c}_j(t)\equiv c_j(t)$. Hence we get from
the application of the lemma for all indices successively that
\begin{equation}\label{sibp}
S=\hbox{$\mit I$\kern-.72em$\mit I$}+{\cal O} (\varepsilon).
\end{equation}
The estimates we are looking for are then just a direct corollary.
\begin{prop}\label{eslf}
Assume there exists a dissipative domain $D_j$ for the index $j$.
Let $\eta_j$ be a loop based at the origin which encircles all the
degeneracies between the real axis and $D_j$ and let $\sigma_j$ be the
permutation of labels associated with $\eta_j$, in the spirit of the
remark ending the previous section. The loop $\eta_j$ is negatively,
respectively
positively, oriented if $D_j$ is above, respectively below, the real axis.
Then the solution of
(\ref{eqco}) subjected to the initial conditions $c_k(-\infty)=
\delta_{jk}$
satisfies
\begin{eqnarray}
c_{\sigma_j(j)}(+\infty)&=&\mbox{e}^{-i\theta_j(\eta_j)}
\mbox{e}^{-i\int_{\eta_j}e_j(z)dz/\varepsilon}\left(1+{\cal O} (\varepsilon)\right)\\
c_{\sigma_j(k)}(+\infty)&=&{\cal O} \left(\varepsilon\mbox{e}^{
\mbox{\em\scriptsize Im }\int_{\eta_j}e_j(z)dz/\varepsilon+h_j
(e_{\sigma_j(j)}(+\infty)-
e_{\sigma_j(k)}(+\infty))/\varepsilon}
\right),
\end{eqnarray}
with $h_j\in [H^-_j,H^+_j]$, where $H^{\pm}_j$ is the maximum,
respectively
minimum, imaginary part of the points at $+\infty$ in $D_j$
\begin{equation}
H^{+}=\lim_{t\rightarrow +\infty}\sup_{s|t+is\in D_j}s,\;\;\;
H^{-}=\lim_{t\rightarrow +\infty}\inf_{s|t+is\in D_j}s.
\end{equation}
\end{prop}
Thus we see that it is possible to get the (exponentially small)
asymptotic
behavior of the element $s_{\sigma_j(j),j}$ of the $S$-matrix, provided
there exists a dissipative domain for the index $j$. The difficult part
of the
problem is of course to prove the existence of such domains $D_j$, which
do not
necessarily exist, and to have enough of them to compute the asymptotic
of
the whole $S$-matrix.
This task is the equivalent for $n$-level systems to the study of the
global
behavior of the Stokes lines for $2$-level systems. We postpone this
aspect
of the problem to the next section. Note that we also get from this
result an
exponential bound on the elements $s_{\sigma_j(k),j}$ of the $S$-matrix,
$k\neq j$, which may or may not be useful. If $\eta_j$ encircles
no point of $\Omega$, we cannot get the asymptotic behavior of
$s_{\sigma_j(j),j}$ but we only get the exponential bounds. Since our
main
concern is asymptotic behaviors, we call the corresponding dissipative
domain
trivial.
\\ {\bf Remark:}\\
In contrast with the $2$-level case, see \cite{jp3}, we have to work
with dissipative
domains instead of working with one dissipative path for all indices.
Indeed, it is not difficult to convince oneself
with specific $3$-level cases that such a dissipative path may not
exist, even
when the eigenvalue degeneracies are close to the real axis. In return,
we prove below
the existence of dissipative domains in this situation.
\\
{\bf Proof:}\\
The asymptotic relation is a direct consequence of lemma \ref{ECHCO},
(\ref{unili}), (\ref{defx}) and the first part of the lemma. The estimate
is a
consequence of the same equations, the second estimate of the lemma and
the
identity, for $t>T$,
\begin{eqnarray}
\mbox{Im }\widetilde{\Delta}_{jk}(t+is)&=&\mbox{Im }\left(\int_{\eta_j}
e_j(z)dz
-\int_{\eta_j}e_k(z)dz\right)\nonumber\\
&+&\int_0^s\mbox{Re }(e_{\sigma_j(j)}(t+is')-
e_{\sigma_j(k)}(t+is'))ds'.
\end{eqnarray}
The path of integration from $0$ to $z$ for $\widetilde{\Delta}_{jk}(z)$
is deformed
into the loop $\eta_j$ followed by the
real axis from $0$ to $\mbox{Re }z$ and a vertical path from
$\mbox{Re }z$
to $z$.
It remains to take the limit $t\rightarrow +\infty$.\hfill {$\Box$}\\
{\bf Proof of lemma \ref{wkb}:}\\
We rewrite equations (\ref{aneqc}) and (\ref{coin}) as an integral
equation and
perform an integration by parts on the exponentials
\begin{eqnarray}
\widetilde{c}_k(z)&=&\delta_{jk}-i\varepsilon\sum_{l=1}^n
\frac{\widetilde{a}_{kl}(z)}{\widetilde{e}_k(z)-\widetilde{e}_l(z)}
\mbox{e}^{i\widetilde{\Delta}_{kl}(z)/\varepsilon}\widetilde{c}_l(z)\nonumber\\
&+&i\varepsilon\sum_{l=1}^n\int_{-\infty}^z
{\left(\frac{\widetilde{a}_{kl}(z')}{\widetilde{e}_k(z')-
\widetilde{e}_l(z')}\right)}'
\mbox{e}^{i\widetilde{\Delta}_{kl}(z')/\varepsilon}\widetilde{c}_l(z')dz'\nonumber\\
&+&i\varepsilon\sum_{l,m=1}^n\int_{-\infty}^z
\frac{\widetilde{a}_{kl}(z')\widetilde{a}_{lm}(z')}
{\widetilde{e}_k(z')-\widetilde{e}_l(z')}
\mbox{e}^{i\widetilde{\Delta}_{km}(z')/\varepsilon}\widetilde{c}_m(z')dz'.
\end{eqnarray}
Since all eigenvalues are distinct in $S_{\alpha}\backslash \Omega$, the
denominators are
always different from $0$. Due to equation (\ref{unili}), the height
above or below the
real axis
at which we start the integration is irrelevant, so that we can use the
symbol $-\infty$ as
lower integration bound. Note that the integrated term vanishes at
$-\infty$. We have
also used the identity
\begin{equation}
\widetilde{\Delta}_{kl}(z')+\widetilde{\Delta}_{lm}(z')
\equiv\widetilde{\Delta}_{km}(z').
\end{equation}
In terms of the functions $x_k$ we get, using the same identity,
\begin{eqnarray}\label{iefx}
x_k(z)&=&\delta_{jk}-i\varepsilon\sum_{l=1}^n
\frac{\widetilde{a}_{kl}(z)}{\widetilde{e}_k(z)-\widetilde{e}_l(z)}
x_l(z)\nonumber\\
&+&i\varepsilon\sum_{l=1}^n\int_{-\infty}^z
{\left(\frac{\widetilde{a}_{kl}(z')}{\widetilde{e}_k(z')-
\widetilde{e}_l(z')}\right)}'
\mbox{e}^{i(\widetilde{\Delta}_{jk}(z)-\widetilde{\Delta}_{jk}(z'))/\varepsilon}
x_l(z')dz'\nonumber\\
&+&i\varepsilon\sum_{l,m=1}^n\int_{-\infty}^z
\frac{\widetilde{a}_{kl}(z')\widetilde{a}_{lm}(z')}{\widetilde{e}_k(z')
-\widetilde{e}_l(z')}
\mbox{e}^{i(\widetilde{\Delta}_{jk}(z)-\widetilde{\Delta}_{jk}(z'))/\varepsilon}
x_m(z')dz'.
\end{eqnarray}
We introduce the quantity
\begin{equation}
|||x|||_j=\sup_{z\in D_j\atop l=1,\cdots,n}|x_l(z)|
\end{equation}
and consider for each $k$ the equation (\ref{iefx}) along the dissipative
path $\gamma^k(u)$ described
in the definition of $D_j$, such that
\begin{equation}
\left|\mbox{e}^{i(\widetilde{\Delta}_{jk}(\gamma^k(t))-
\widetilde{\Delta}_{jk}(\gamma^k(u)))/\varepsilon}\right|\leq 1
\end{equation}
when $u\leq t$ along that path. Due to the integrability of the
$\widetilde{a}_{kl}(z)$ at infinity and the uniform boundedness of
$d\gamma^k(u)/du$, we get the estimate
\begin{equation}
|x_k(z)-\delta_{kj}|\leq \varepsilon |||x|||_j A
\end{equation}
for some constant $A$ uniform in $z\in D_j$, hence
\begin{equation}
|||x|||_j\leq 1+\varepsilon |||x|||_j A.
\end{equation}
Consequently, for $\varepsilon$ small enough
\begin{equation}
|||x|||_j\leq 2.
\end{equation}
And the result follows.\hfill {$\Box$}
\section{Superasymptotic improvement}\label{super}
\setcounter{equation}{0}
All results above can be substantially improved by using the so-called
superasymptotic renormalization method \cite{be}, \cite{n}, \cite{jp5}.
The joint use of complex WKB analysis
and superasymptotic renormalization is very powerful, as demonstrated
recently
in \cite{jp3} for $2$-level systems, and allows, roughly speaking, to
replace all
remainders
${\cal O}(\varepsilon)$ by ${\cal O} (\mbox{e}^{-\kappa /\varepsilon})$, where $\kappa >0$. We
briefly
show how to achieve this improvement in the case of $n$-level systems.
Let $H(z)$ satisfy H1, H2 and H3 in $S_{\alpha}$ and let
\begin{equation}
\widehat{S}_{\alpha}=S_{\alpha}\backslash \cup_{r=1,\cdots,p}
(J_r\cup\overline{J_r}),
\end{equation}
where each $J_r$ is an open domain containing one point of $\Omega$ only
in the
open upper half plane. Hence, any analytic continuation $e_j(z)$ of
$e_j(t)$,
$t\in\hbox{$\mit I$\kern-.277em$\mit R$}$, in
$\widehat{S}_{\alpha}$ is isolated in the spectrum of $H(z)$ so that
$e_j(z)$ is analytic and multivalued in $\widehat{S}_{\alpha}$, and the
same
is true for the corresponding analytic continuation $P_j(z)$ of
$P_j(t)$, $t\in\hbox{$\mit I$\kern-.277em$\mit R$}$. Let $\sigma_r$ be the permutation associated with
the loop $\zeta_r$ based at the origin which encircles $J_r$ once, such
that
\begin{equation}\label{mojr}
\widetilde{e}_j(z)={e}_{\sigma_r(j)}(z),
\end{equation}
with the convention of section \ref{anprop}. The matrix $K(z)$ is
analytic
and single valued in $\widehat{S}_{\alpha}$. Consider the single valued
analytic matrix
\begin{equation}
H_1(z,\varepsilon)=H(z)-i\varepsilon K(z),\;\;\; z\in \widehat{S}_{\alpha}.
\end{equation}
For $\varepsilon$ small enough, the spectrum of $H_1(z,\varepsilon)$ is non degenerate
$\forall z\in \widehat{S}_{\alpha}$ so that its eigenvalues
$e_j^1(z,\varepsilon)$ and
eigenprojectors $P_j^1(z,\varepsilon)$ are multivalued analytic functions in
$\widehat{S}_{\alpha}$. Moreover, for $\varepsilon$ small enough, the analytic
continuations of $e_j^1(z,\varepsilon)$ and $P_j^1(z,\varepsilon)$ around $J_r$
satisfy
\begin{eqnarray}
\widetilde{e}^1_j(z)&=&{e}^1_{\sigma_r(j)}(z)\\
\widetilde{P}^1_j(z)&=&{P}^1_{\sigma_r(j)}(z),
\end{eqnarray}
as can be easily deduced from (\ref{mojr}) by perturbation theory.
Consequently the
matrix
\begin{equation}
K_1(z,\varepsilon)=\sum_{j=1}^m{{P}^1_j}'(z,\varepsilon){P}^1_j(z,\varepsilon)
\end{equation}
is analytic and single valued in $\widehat{S}_{\alpha}$. Defining the
single
valued matrix
\begin{equation}
H_2(z,\varepsilon)=H(z)-i\varepsilon K_1(z,\varepsilon),\;\;\; z\in \widehat{S}_{\alpha},
\end{equation}
we can repeat the argument, for $\varepsilon$ small enough. By induction we set
for
any $q\in\hbox{$\mit I$\kern-.277em$\mit N$}$,
\begin{eqnarray}
H_q(z,\varepsilon)&=&H(z)-i\varepsilon K_{q-1}(z,\varepsilon)\\
K_{q-1}(z,\varepsilon)&=&\sum_{j=1}^m{{P}^{q-1}_j}'(z,\varepsilon){P}^{q-1}_j
(z,\varepsilon),
\;\;\; z\in \widehat{S}_{\alpha}
\end{eqnarray}
for $\varepsilon$ is small enough. We have
\begin{equation}
H_q(z,\varepsilon)=\sum_{j=1}^m e^{q}_j(z,\varepsilon){P}^{q}_j(z,\varepsilon),
\end{equation}
where the eigenvalues and eigenprojections are multivalued in
$\widehat{S}_{\alpha}$ and satisfy
\begin{eqnarray}
\widetilde{e}^{q}_j(z,\varepsilon)&=&e^{q}_{\sigma_r(j)}(z,\varepsilon)\\
\widetilde{P}^{q}_j(z,\varepsilon)&=&P^{q}_{\sigma_r(j)}(z,\varepsilon),\;\;\; j=1,
\cdots,n,
\end{eqnarray}
with the notations of (\ref{mojr}). We quote from \cite{jp3}, \cite{jp5}
the
main proposition regarding this construction.
\begin{prop}\label{supa}
Let $H(z)$ satisfy H1, H2 and H3 in $S_{\alpha}$ and let
$\widehat{S}_{\alpha}$ be defined as above. Then there exist constants
$c>0$, $\varepsilon^*>0$ and a real function $b(t)$ with
$\lim_{t\rightarrow\pm\infty}|t|^{1+a}b(t)<\infty$, such that
\begin{eqnarray}
& &\|K_q(z,\varepsilon)-K_{q-1}(z,\varepsilon)\|\leq b(\mbox{\em Re }z)
\varepsilon^qc^qq!\\
& &\|K_q(z,\varepsilon)\|\leq b(\mbox{\em Re }z),
\end{eqnarray}
for all
$z\in\widehat{S}_{\alpha}$, all $\varepsilon<\varepsilon^*$ and all
$q\leq q^*(\varepsilon)\equiv\left[1/ec\varepsilon\right]$, where $[y]$ denotes the
integer part of $y$ and $e$ is the basis of the neperian logarithm.
\end{prop}
We can deduce from this that in $\widehat{S}_{\alpha}$
\begin{eqnarray}\label{peq}
e_j^q(z,\varepsilon)&=&e_j(z)+{\cal O} (\varepsilon^2b(\mbox{Re }z))\\
P_j^q(z,\varepsilon)&=&e_j(z)+{\cal O} (\varepsilon(\mbox{Re }z)),\;\;\;\forall
q\leq q^*(\varepsilon).
\end{eqnarray}
We introduce the notation $f^{q^*(\varepsilon)}\equiv f^*$ for any quantity
$f^q$ depending on the index $q$ and we drop from now on the $\varepsilon$ in
the arguments of the functions we encounter. We define the multivalued
analytic matrix $W_*(z)$ for $z\in \widehat{S}_{\alpha}$ by
\begin{equation}
{W_*}'(z)=K_*(z)W_*(z), \;\;\; W_*(0)=\hbox{$\mit I$\kern-.72em$\mit I$}.
\end{equation}
Due to the above observations and proposition \ref{supa}, $W_*(z)$ enjoys
all properties $W(z)$ does, such as
\begin{eqnarray}
& &W_*(z)P_j^*(z)=P_j^*(0)W_*(z)\\
& &\widetilde{W}^*(z)=W_*(z)W_*(\zeta_r)
\end{eqnarray}
and, uniformly in $s$,
\begin{equation}
\lim_{t\pm\infty}W_*(t+is)=W_*(\infty).
\end{equation}
Thus we define for any $z\in \widehat{S}_{\alpha}$ as set of
eigenvectors of
$H_*(z)$ by
\begin{equation}
\varphi_j^*(z)= W_*(z) \varphi_j^*(0),
\end{equation}
where
\begin{equation}
H_*(0)\varphi_j^*(0)=e^*_j(0)\varphi_j^*(0), \;\;\; j=1,\cdots, n,
\end{equation}
and which satisfy
\begin{equation}
\widetilde{\varphi}_j^*(0)=\mbox{e}^{-i\theta_j^*(\zeta_r)}
\varphi_{\sigma_r(j)}^*(0),
\end{equation}
with
\begin{equation}
\theta_j^*(\zeta_r)=\theta(\zeta_r)+{\cal O} (\varepsilon)\in \hbox{$\mit I$\kern-.7em$\mit C$}.
\end{equation}
Let us expand the solution of (\ref{schr}) on this multivalued set of
eigenvectors as
\begin{equation}\label{decsup}
\psi(z)=\sum_{j=1}^nc^*_j(z)\mbox{e}^{-i\int_0^ze^*_j(z')dz'/\varepsilon}\varphi^*_j(z).
\end{equation}
Since the analyticity properties of the eigenvectors and eigenvalues of
$H_*(z)$ are
the same as those enjoyed by the eigenvectors and eigenvalues of $H(z)$,
we get
as in lemma \ref{ECHCO}
\begin{equation}
\widetilde{c}_j^*(z)\mbox{e}^{-i\int_{\zeta_r}e_j^*(u)du/\varepsilon}
\mbox{e}^{-i\theta_j^*(\zeta_r)}
=c_{\sigma_r(j)}^*(z), \;\;\; \forall z\in \widehat{S}_{\alpha}.
\end{equation}
Substituting (\ref{decsup}) in (\ref{schr}), we see that the multivalued
coefficients
${c}_j^*(z)$ satisfy in $\widehat{S}_{\alpha}$ the differential equation
\begin{equation}
{c_j^*}'(z)=\sum_{k=1}^na_{jk}^*(z)\mbox{e}^{i\Delta_{jk}^*(z)/\varepsilon}c_k^*(z)
\end{equation}
where
\begin{equation}
\Delta_{jk}^*(z)=\int_0^ze_j^*(z')-e_k^*(z')dz'
\end{equation}
and
\begin{equation}
a_{jk}^*(z)=\frac{\langle\varphi_j^*(z)(0)|P_j^*(z)(0){W_*(z)}^{-1}
(K_{q^*-1}(z)-K_{q^*}(z))W_*(z)\varphi_k^*(0)\rangle}{\|\varphi_j^*(0)\|^2},
\end{equation}
to be compared with (\ref{anajk}).
The key point of this construction is that it follows from proposition
\ref{supa} with $q=q^*(\varepsilon)$ that
\begin{equation}\label{key}
|a_{jk}^*(z)|\leq 2 b(\mbox{Re }z)\mbox{e}^{-\kappa/\varepsilon},\;\;\; \forall
z\in \widehat{S}_{\alpha}
\end{equation}
where $\kappa=1/ec >0$, and from (\ref{peq}) that
\begin{equation}\label{keyd}
\mbox{Im }\Delta_{jk}^*(z)=\mbox{Im }\Delta_{jk}(z)+{\cal O} (\varepsilon^2),
\end{equation}
uniformly in $z\in \widehat{S}_{\alpha}$. Thus, we deduce from
(\ref{key})
that the limits
\begin{equation}
\lim_{t\rightarrow\pm\infty}c_j^*(t+is)=c_j^*(\pm\infty), \;\;\; j=1,\cdots,n,
\end{equation}
exist for any analytic continuation in $\widehat{S}_{\alpha}$. Moreover,
along any dissipative path $\gamma^k(u)$ for $\{jk\}$, as defined above,
we get from
(\ref{keyd})
\begin{equation}
\left|\mbox{e}^{i(\widetilde{\Delta}_{jk}^*(\gamma^k(t))-
\widetilde{\Delta}_{jk}^*(\gamma^k(u)))/\varepsilon}\right|={\cal O} (1)
\end{equation}
so that, reproducing the proof of lemma \ref{wkb} we have
\begin{lem}
In a dissipative domain $D_j$, if
$\widetilde{c}_k^*(-\infty)={c}_k^*(-\infty)=\delta_{kj}$, then
\begin{eqnarray}
& &\widetilde{c}_j^*(z)=1+{\cal O} (\mbox{e}^{-\kappa/\varepsilon})\\
& &\mbox{e}^{i\widetilde{\Delta}_{jk}(z)\varepsilon}\widetilde{c}_k^*(z)={\cal O}
\left(\mbox{e}^{-\kappa/\varepsilon}\right)
,\;\;\;\forall k\neq j,
\end{eqnarray}
uniformly in $z\in \widehat{S}_{\alpha}$.
\end{lem}
This lemma
yields the improved version of our main result.
\begin{prop}\label{MAPROP}
Under the conditions of proposition \ref{eslf} and with the
same notations. If $c_k^*(-\infty)=\delta_{jk}$, then
\begin{eqnarray}
c_{\sigma_j(j)}^*(+\infty)&=&\mbox{e}^{-i\theta_j^*(\eta_j)}
\mbox{e}^{-i\int_{\eta_j}e_j^*(z)dz/\varepsilon}\left(1+{\cal O} \left(\mbox{e}^{-\kappa/\varepsilon}
\right)\right)\\
c_{\sigma_j(k)}^*(+\infty)&=&{\cal O} \left(\mbox{e}^{-\kappa/\varepsilon}\mbox{e}^{
\mbox{\em\scriptsize Im }\int_{\eta_j}e_j(z)dz/\varepsilon+h_j(e_{\sigma_j(j)}
(+\infty)-
e_{\sigma_j(k)}(+\infty))/\varepsilon}
\right).
\end{eqnarray}
\end{prop}
Note that we may or may not replace $e_j(z)$ by $e_j^*(z)$ in the
estimate
without altering the result.
It remains to make the link between the $S$-matrix and the
$c_k^*(+\infty)$'s
of the proposition explicit. We define $\beta_j^{*\pm}$ by the
relations
($H_*(z)$ and $H(z)$ coincide at $\pm\infty$),
\begin{equation}
\varphi_j^*(\pm\infty)=\mbox{e}^{-i\beta_j^{*\pm}}\varphi_j(\pm\infty).
\end{equation}
By comparison of (\ref{decsup}) and (\ref{expa}) we deduce the lemma
\begin{lem}\label{sdec}
If $c_k(t)$ and $c_k^*(t)$ satisfy $c_k(-\infty)=c_k^*(-\infty)=
\delta_{jk}$,
then, the element $kj$ of the $S$-matrix is given by
\begin{eqnarray}
s_{kj}&=&c_k(+\infty)=\mbox{e}^{-i(\beta_k^{*+}-\beta_j^{*-})}
\mbox{e}^{-i\int_0^{+\infty}
e_k^*(t')-e_k(t')dt'/\varepsilon}\mbox{e}^{-i\int_{-\infty}^0e_j^*(t')-
e_j(t')dt'/\varepsilon}
c_k^*(+\infty)\nonumber\\
&\equiv& \mbox{e}^{-i\alpha_{kj}^*}c_k^*(+\infty),
\end{eqnarray}
with $\beta_j^{*\pm}={\cal O} (\varepsilon)$ and
$\int_{\pm\infty}^0e_j^*(t')-e_j(t')dt'/\varepsilon={\cal O} (\varepsilon)$, i.e.
$\mbox{e}^{-i\alpha_{kj}^*}=1+{\cal O} (\varepsilon)$.
\end{lem}
{\bf Remarks:}
\\ i) Proposition \ref{MAPROP} together with lemma \ref{sdec} are the
main results of the
first part of this paper.
\\ ii) As a direct consequence of these estimates on the real axis we
have
\begin{equation}\label{sjksup}
s_{jk}={\cal O} \left(\mbox{e}^{-\kappa/\varepsilon}\right),\;\;\; \forall k\neq j,
\end{equation}
and
\begin{equation}
s_{jj}=\mbox{e}^{-i\alpha_{jj}^*}\left(1+{\cal O} \left(e^{-\kappa /\varepsilon}\right)
\right).
\end{equation}
iii) It should be clear from the analysis just performed that all results
obtained hold if the generator $H(z)$ in (\ref{schr}) is replaced by
\begin{equation}
H(z,\varepsilon)=H_0(z)+{\cal O} (\varepsilon b(\mbox{Re }z)),
\end{equation}
with $b(t)={\cal O} (1/t^{1+a})$,
provided $H_0(z)$ satisfies the hypotheses we assumed.
\section{Avoided crossings}\label{avoid}
\setcounter{equation}{0}
We now come to the second part of the paper in which we prove asymptotic
formulas for
the off-diagonal elements of the $S$-matrix, by means of the general set
up presented
above.
To start with, we define a class of $n$-level systems for which we can
prove the
existence of one non trivial dissipative domains for all indices.
They are obtained by means
of systems exhibiting degeneracies of eigenvalues on the real axis,
hereafter
called real crossings, which we
perturb in such a way that these degeneracies are lifted and turn to
avoided
crossings on the real axis. When the perturbation is small enough, this
process
moves the eigenvalue degeneracies off the real axis but they
remain close to the place where the real crossings occurred.
This method was used successfully in \cite{j1} to deal with $2$-level
systems.
We do not attempt to list
all cases in which dissipative domains can be constructed by means of
this technique
but rather present a wide class of examples which are relevant in the
theory of quantum adiabatic
transitions and in the theory of multichannel semiclassical scattering,
as described
below.
\\
Let $H(t,\delta)\in M_n(\hbox{$\mit I$\kern-.7em$\mit C$})$ satisfy the following assumptions.
\\
{\bf H4} {For each fixed $\delta\in[0,d]$, the matrix
$H(t,\delta)$ satisfies H1 in a strip $S_{\alpha}$ independent of
$\delta$ and $H(z,\delta)$,
$\partial /\partial z H(z,\delta)$ are
continuous as a functions of two variables $(z,\delta)\in S_{\alpha}
\times [0,d]$.
Moreover, it satisfies H2 uniformly in
$\delta\in[0,d]$, with limiting values $H(\pm,\delta)$ which are
continuous
functions of $\delta\in [0,d]$.}\\
{\bf H5} {\em For each $t\in\hbox{$\mit I$\kern-.277em$\mit R$}$ and each $\delta\in[0,d]$, the spectrum
of $H(t,\delta)$, denoted
by $\sigma(t,\delta)$, consists in $n$ real eigenvalues
\begin{equation}
\sigma(t,\delta)=\{e_1(t,\delta), e_2(t,\delta),\cdots,e_n(t,\delta)\}
\subset \hbox{$\mit I$\kern-.277em$\mit R$}
\end{equation}
which are distinct when $\delta>0$
\begin{equation}
e_1(t,\delta)< e_2(t,\delta)<\cdots <e_n(t,\delta).
\end{equation}
When $\delta =0$, the functions $e_j(t,0)$ are analytic on the real axis
and
there exists a finite set
of crossing points $\{t_1\leq t_2\leq \cdots \leq t_p\}\in\hbox{$\mit I$\kern-.277em$\mit R$}$, $p\geq 0$,
such that
\\
i) $\forall t<t_1$,
\begin{equation}
e_1(t,0)< e_2(t,0)<\cdots <e_n(t,0).
\end{equation}
ii) $\forall j<k\in\{1,2,\cdots ,n\}$, there exists at most one $t_r$
with
\begin{equation}
e_j(t_r,0)-e_k(t_r,0)=0,
\end{equation}
and if such a $t_r$ exists we have
\begin{equation}
\frac{\partial}{\partial t}\left(e_j(t_r,0)-e_k(t_r,0)\right)>0.
\end{equation}
iii) $\forall j\in\{1,2,\cdots ,n\}$, the eigenvalue $e_j(t,0)$ crosses
eigenvalues whose
indices are all superior to $j$ or all inferior to $j$.
}
\noindent{\bf Remarks:}
\\ i) The parameter $\delta$ can be understood as a coupling constant
controlling
the strength of the perturbation.
\\ ii)The eigenvalues $e_j(t,0)$ are assumed to be analytic on the real
axis,
because of the degeneracies on the real axis. However, if $H(t,\delta)$
is self adjoint
for any $\delta\in[0,d]$, this follows from a theorem of Rellich, see
\cite{k}.
\\ iii) We give in figure \ref{figlev} an example of pattern of crossings
with
the corresponding pattern of avoided crossings for which the above
conditions are fulfilled.
\\ iv) The crossings are assumed to be generic
in the sense that the derivative of $e_j-e_k$ are non zero at the
crossing $t_r$.
\\ v) The crossing points $\{t_1,t_2,\cdots ,t_p\}$ need not be distinct,
which is important when
the eigenvalues possess symmetries. However, for each $j=1,\cdots, n$,
the eigenvalue
$e_j(t,\delta)$ experiences avoided crossings with $e_{j+1}(t,\delta)$
and/or
$e_{j-1}(t,\delta)$ at a subset of distinct points
$\{t_{r_1},\cdots ,t_{r_j}\}\subseteq \{t_1,t_2,\cdots ,t_p\}$.
\begin{figure}
\vspace{7cm}
\hspace{3cm}\special{picture fig4 scaled 500}
\caption{A pattern of eigenvalue crossings (bold curves) with
the corresponding pattern of avoided crossings (fine curves)
satisfying H5.}\label{figlev}
\end{figure}
We now state the main lemma of this section regarding the analyticity
properties of the perturbed levels and the existence of
dissipative domains for all indices in this perturbative context.
\begin{lem}\label{MAINL}
Let $H(t,\delta)$ satisfy H4 and H5. We can choose $\alpha>0$ small
enough so
that the following assertions are true for sufficiently small
$\delta>0$:
\\ i) Let $\{t_{r_1},\cdots ,t_{r_j}\}$ be the set of avoided crossing
points experienced by $e_j(t,\delta)$, $j=1,\cdots, n$.
For each $j$, there exists a
set of distinct domains $J_r\in S_{\alpha}$, where
$r\in \{{r_1},\cdots ,{r_j}\}$,
\begin{equation}
J_r=\{z=t+is|\,0\leq |t-t_r|<L,\, 0<g<s<\alpha'\},
\end{equation}
with $L$ small enough, $\alpha'<\alpha$ and $g>0$ such that
$e_j(-\infty,\delta)$ can be analytically continued in
\begin{equation}
S^j_{\alpha}=S_{\alpha}\backslash \cup_{r=r_1,\cdots,r_j}
(J_r\cup\overline{J_r}).
\end{equation}
ii) Let $t_r$ be an avoided crossing point of $e_j(t,\delta)$ with
$e_{k}(t,\delta)$, $k=j\pm 1$.
Then the analytic continuation of $e_j(t_r,\delta)$ along a loop based
at $t_r\in\hbox{$\mit I$\kern-.277em$\mit R$}$ which encircles $J_r$ once yields
$\widetilde{e}_j(t_r,\delta)$ back at $t_r$ with
\begin{equation}
\widetilde{e}_j(t_r,\delta)=e_{k}(t_r,\delta).
\end{equation}
iii) For each $j=1,\cdots, n$,
there exists a dissipative domain $D_j$ above or below the real axis in
$S_{\alpha}\cap\{z=t+is|\,|s|\geq \alpha'\}$.
The permutation $\sigma_j$ associated with these
dissipative domains (see proposition \ref{eslf}) are all given by
$\sigma_j=\sigma$
where $\sigma$ is the permutation which maps the index
of the $k^{\scriptsize \mbox{th}}$ eigenvalue $e_j(\infty,0)$ numbered
from the lowest one on $k$, for all $k\in\{1,2,\cdots ,n\}$.
\end{lem}
{\bf Remarks:}
\\ i) In part ii) the same result is true along a loop encircling
$\overline{J_r}$.
\\ ii) The dissipative domains $D_j$ of part iii) are located above
(respectively below) all the sets $J_r$ (resp. $\overline{J_r}$),
$r=1,\cdots, p$.
\\ iii) The main interest of this lemma is that the sufficient
conditions
required for the existence of dissipative domains in the complex plane
can be deduced from the behavior of the eigenvalues on the {\em real}
axis.
\\ iv) We emphasize that more general types of avoided crossings than
those described in
H5 may lead to the existence of dissipative domains for {\em certain}
indices
but we want to get dissipative domains for {\em all} indices. For
example, if
part iii) of H5 is satisfied for certain indices only, then part iii) of
lemma \ref{MAINL}
is satisfied for those indices only.
\\ v) Note also that there are patterns of eigenvalue crossings for which
there
exist no dissipative domain for some indices. For example, if $e_j(t,0)$
and $e_k(t,0)$
display two crossings, it is not difficult to see from the proof of the
lemma
that no dissipative domains can exist for $j$ or $k$.
We postpone the proof of this lemma to the end of the section and go on
with
its consequences.
By applying the results of the previous section we get
\begin{thm}\label{PERCO}
Let $H(t,\delta)$ satisfy H4 and H5. If $\delta>0$ is small enough,
the
elements $\sigma(j)j$ of the $S$-matrix, with $\sigma(j)$ defined in
lemma \ref{MAINL}, are given in the limit $\varepsilon\rightarrow 0$ for all $j=1,
\cdots, n$, by
\begin{equation}
s_{\sigma(j)j}=\prod_{k=j}^{\sigma(j)\mp1}\mbox{e}^{-i\theta_k(\zeta_k)}
\mbox{e}^{-i\int_{\zeta_k}
e_k(z,\delta)dz/\varepsilon}\left(1+{\cal O} (\varepsilon)\right),\;\;\;\sigma(j)
\left\{{ >j \atop <j}\right.
\end{equation}
where, for $\sigma(j)>j$ (respectively $\sigma(j)<j$), $\zeta_k$,
$k=j,\cdots,\sigma(j)\mp 1$, denotes a
negatively (resp. positively) oriented loop
based at the origin which encircles the set $J_r$ (resp.
$\overline{J_r}$
corresponding to the avoided crossing between $e_k(t,\delta)$ and
$e_{k+1}(t,\delta)$ (resp. $e_{k-1}(z,\delta)$) at $t_r$,
$\int_{\zeta_k}e_k(z,\delta)dz$ denotes the integral along $\zeta_k$
of the
analytic continuation of $e_k(0,\delta)$
and $\theta_k(\zeta_k)$
is the corresponding factor defined by (\ref{theta}), see figure
\ref{ficor}.
More accurately, with the notations of section \ref{super}, we have
the improved formula
\begin{equation}\label{imfor}
s_{\sigma(j)j}=\mbox{e}^{-i\alpha_{\sigma(j)j}^*}
\prod_{k=j}^{\sigma(j)\mp1}\mbox{e}^{-i\theta_k^*(\zeta_k)}
\mbox{e}^{-i\int_{\zeta_k}
e_k^*(z,\delta)dz/\varepsilon}\left(1+{\cal O} \left(e^{-\kappa /\varepsilon}\right)
\right),
\;\;\;\sigma(j)\left\{{ >j \atop <j}\right.
\end{equation}
The element $\sigma(l)j$, $l\neq j$, are estimated by
\begin{equation}
s_{\sigma(l)j}={\cal O} \left(\varepsilon \mbox{e}^{h(e_{\sigma(j)}(\infty,\delta)-
e_{\sigma(l)}(\infty,\delta))/\varepsilon}\prod_{k=j}^{\sigma(j)\mp1}\mbox{e}^{
\mbox{\em \scriptsize Im }\int_{\zeta_k}
e_k(z,\delta)dz/\varepsilon} \right),\;\;\;\sigma(j)\left\{{ >j \atop <j}
\right.
\end{equation}
where $h$ is strictly positive (resp. negative) for $\sigma(j)>j$
(respectively $\sigma(j)<j$).
\end{thm}
{\bf Remarks:}
\\ i) As the eigenvalues are continuous at the degeneracy points, we
have that
\begin{equation}
\lim_{\delta\rightarrow 0}\mbox{e}^{-i\int_{\zeta_k}e_k(z,\delta)dz}=0,\;\;\;\forall
k=1,\cdots, p.
\end{equation}
ii) The remainders ${\cal O}(\varepsilon)$ depend on $\delta$ but it should be
possible to get
estimates which are valid as both $\varepsilon$ and $\delta$ tend to zero, in
the spirit of
\cite{j1}, \cite{mn} and \cite{r}.
\\ iii) This result shows that one off-diagonal element per column of
the $S$-matrix at least can be
computed asymptotically. However, it is often possible to get more
elements by making use
of symmetries of the $S$-matrix.
Moreover, if there exist dissipative domains going
above or below other eigenvalue degeneracies further away in the complex
plane, other
elements of the $S$-matrix can be computed.
\\ iv) Finally, note that all starred quantities in (\ref{imfor})
depend on $\varepsilon$.
\begin{figure}
\vspace{6cm}
\hspace{3cm}\special{picture fig5 scaled 600}
\caption{The loops $\eta_j$ and $\zeta_k$, $k=j,\cdots,\sigma(j)-1$.}
\label{ficor}
\end{figure}
\\{\bf Proof:} The first thing to determine is whether the loops
$\zeta_k$ are
above or below the real axis. Since the formulas we
deduce from the complex WKB analysis are asymptotic formulas, it suffices
to
choose the case which yields exponential decay of $s_{\sigma(j)j}$. It
is
readily checked in the proof of lemma \ref{MAINL} below that if
$\sigma(j)>j$,
$D_j$ is above the real axis and if $\sigma(j)<j$, $D_j$ is below the
real axis.
Then it
remains to explain how to pass from the loop $\eta_j$ given in
proposition \ref{eslf} to the set of loops $\zeta_k$, $k=j,\cdots,
\sigma(j)-1$.
We briefly deal with the case $\sigma(j)>j$, the other case being
similar.
It follows from lemma \ref{MAINL} that we can deform $\eta_j$ into the
set of
loops $\zeta_k$, each associated with one avoided crossing, as described
in
figure \ref{ficor}. Thus we have
\begin{equation}
\int_{\eta_j}=\sum_{k=j}^{\sigma(j)-1}\int_{\zeta_k}
\end{equation}
for the decay rate and, see \ref{mono},
\begin{equation}
W(\eta_j)=W(\zeta_{\sigma(j)-1})\cdots W(\zeta_{j+1})W(\zeta_j)
\end{equation}
for the prefactors.
Let $\nu_j$ be a negatively oriented loop based at $t_r$
which encircles $J_r$ as described in lemma \ref{MAINL}. Consider now
the loop
$\zeta_j$ associated with
this avoided crossing and deform it to the path obtained by going from
$0$
to $t_r$ along the real axis, from $t_r$ to $t_r$ along $\nu_j$ and
back from $t_r$ to the origin along the real axis. By the point ii) of
the lemma
we get
\begin{equation}
\widetilde{e}_j(0,\delta)={e}_{j+1}(0,\delta)
\end{equation}
along $\zeta_j$, and, accordingly (see (\ref{theta})),
\begin{equation}
\widetilde{\varphi}_j(0,\delta)=\mbox{e}^{-i\theta_j(\zeta_j)}{\varphi}_{j+1}
(0,\delta).
\end{equation}
This justifies the first factor in the formula. By repeating the argument
at the next avoided crossings, keeping in mind that we get
${e}_{j+1}(0,\delta)$
at the end of $\zeta_j$ and so on, we get the final result.
The estimate on $s_{\sigma(l)j}$ is obtained by direct application of
lemma \ref{MAINL}.
\hfill {$\Box$}
\\
{\bf Proof of lemma \ref{MAINL}:}
In the sequel we shall denote "$\frac{\partial}{\partial t}$" by a
"$\prime$ ".
We have to consider the analyticity properties of
$\widetilde{e}_j(z,\delta)$
and define domains in which every point $z$ can be reached
from $-\infty$ by means of a path $\gamma(u)$, $u\in]-\infty,t]$,
$\gamma(t)=z$ such that
$\mbox{Im }\widetilde{\Delta}_{jk}(\gamma(u),\delta)$
is non decreasing in $u$ for certain indices $j\neq k$ when $\delta >0$
is fixed.
Note that by Schwarz's principle if
$\gamma(u)$ is dissipative for $\{jk\}$, then $\overline{\gamma(u)}$ is
dissipative for $\{kj\}$.
When $\gamma(u)=\gamma_1(u)+i\gamma_2(u)$ is differentiable, saying that
$\gamma(u)$
is dissipative for $\{jk\}$ is equivalent to
\begin{eqnarray}\label{disco}
& &\mbox{Re }(\widetilde{e}_j(\gamma(u),\delta)-
\widetilde{e}_k(\gamma(u),\delta))\dot{\gamma}_2(u)
+\mbox{Im }(\widetilde{e}_j(\gamma(u),\delta)-
\widetilde{e}_k(\gamma(u),\delta))\dot{\gamma}_1(u)
\geq 0\nonumber\\
& &\hfill\forall u\in]-\infty,t]
\end{eqnarray}
where "$\dot{\hspace{.2cm}}$" denotes the derivative with respect to $u$.
Moreover, if the eigenvalues are analytic in a neighborhood of the real
axis,
we have the relation in that neighborhood
\begin{equation}\label{imre}
\mbox{Im }(\widetilde{e}_j(t+is,\delta)-
\widetilde{e}_k(t+is,\delta))=\int_0^s\mbox{Re }
(\widetilde{e}_j'(t+is',\delta)-
\widetilde{e}_k'(t+is',\delta))ds',
\end{equation}
which is a consequence of the Cauchy-Riemann identity.
We proceed as follows. We construct dissipative domains above and below
the real
axis when $\delta=0$ and we show that they remain dissipative for the
perturbed quantities $\widetilde{\Delta}_{jk}(z,\delta)$ provided
$\delta$ is small enough.
We introduce some quantities to be used in the construction. Let
$C_r\in \{ 1,\cdots,n\}^2$ denote
the set of distinct couples of indices such that the corresponding
eigenvalues experience
one crossing at $t=t_r$. Similarly, $N\in \{ 1,\cdots,n\}^2$ denotes
the set of couples of indices such that the corresponding eigenvalues
never
cross.
\\Let $I_r=[t_r-L,t_r+L]\in\hbox{$\mit I$\kern-.277em$\mit R$}$, $r=1,\cdots,p$, with $L$ so small that
\begin{equation}
\min_{r\in \{ 1,\cdots,p\}}\min_{\{jk\}\in C_r,\, j<k}\inf_{t\in I_r}
(e_j'(t,0)-e_{k}'(t,0))\equiv 4c>0.
\end{equation}
This relation defines the constant $c$ and we also define $b$ by
\begin{eqnarray}
& &\min_{r\in \{ 1,\cdots,p\}}\min_{\{jk\}\in C_r,\, j<k}
\inf_{t\in \hbox{$\mit I$\kern-.277em$\mit R$}\backslash I_r}
\left|e_j(t,0)-e_{k}(t,0)\right|\geq 4b>0\\
& &\min_{\{jk\}\in N,\, j<k}\inf_{t\in \hbox{$\mit I$\kern-.277em$\mit R$}}\left|e_j(t,0)-e_{k}(t,0)
\right|\geq 4b>0.
\end{eqnarray}
We further introduce
\begin{equation}
I_r^{\alpha}=\{z=t+is|t\in I_r, |s|\leq \alpha\},\;\;\; r=1,\cdots, p.
\end{equation}
Then we choose $\alpha$ small enough so that the only points of
degeneracy of eigenvalues
in $S_{\alpha}$ are on the real axis and
\begin{eqnarray}
& &\label{dec}
\min_{r\in \{ 1,\cdots,p\}}\min_{\{jk\}\in C_r,\, j<k}
\inf_{z\in I_r^{\alpha}}
\mbox{Re }(e_j'(z,0)-e_{k}'(z,0))> 2c>0\\
& &\min_{r\in \{ 1,\cdots,p\}}\min_{\{jk\}\in C_r,\, j<k}
\inf_{z\in S_{\alpha}
\backslash I_r^{\alpha}}\left|
\mbox{Re }(e_j(z,0)-e_{k}(z,0))\right|> 2b>0\\
& &\min_{\{jk\}\in N,\, j<k}\inf_{z\in S_{\alpha}}\left|
\mbox{Re }(e_j(z,0)-e_{k}(z,0))\right|> 2b>0.
\end{eqnarray}
That this choice is always possible is a consequence of the analyticity
of $e_j(z,0)$ close to
the real axis and of the fact that we can work essentially in a compact,
because of hypothesis H4. Let $a(t)$ be integrable on $\hbox{$\mit I$\kern-.277em$\mit R$}$ and such
that
\begin{equation}\label{defa}
\frac{a(t)}{2}>\max_{j< k\in\{ 1,\cdots,n\}}\sup_{|s|\leq\alpha}
\left|\mbox{Re }
(e_j'(t+is,0)-e_{k}'(t+is,0))\right|.
\end{equation}
It follows from H4 that such functions exist.
Let $r\in\{1,\cdots,p\}$ and $\gamma_2(u)$ be a solution of
\begin{equation}
\left\{\matrix{\dot{\gamma}_2(u)=-\frac{\gamma_2(u)a(u)}{b}
\hfill& u\in]-\infty,t_r-L]\hfill\cr
\dot{\gamma}_2(u)=0 \hfill& u\in]t_r-L,t_r+L[\hfill\cr
\dot{\gamma}_2(u)=+\frac{\gamma_2(u)a(u)}{b}\hfill& u
\in[t_r+L,\infty[\hfill}\right.
\end{equation}
with $\gamma_2(t_r)>0$. Then $\gamma_2(u)>0$ for any $u$, since
\begin{equation}
\left\{\matrix{{\gamma}_2(u)=\gamma_2(t_r)\mbox{e}^{-\int_{t_r-L}^u
a(u')du'/b}
\hfill& u\in]-\infty,t_r-L]\hfill\cr
{\gamma}_2(u)=\gamma_2(t_r) \hfill& u\in]t_r-L,t_r+L[
\hfill\cr
{\gamma}_2(u)=\gamma_2(t_r)\mbox{e}^{\int_{t_r+L}^ua(u')du'/b}
\hfill& u\in[t_r+L,\infty[\hfill}\right.
\end{equation}
and since $a(u)$ is integrable, the limits
\begin{equation}
\lim_{u\rightarrow\pm\infty}\gamma_2(u)=\gamma_2(\pm\infty)
\end{equation}
exist. Moreover, we can always choose $\gamma_2(t_r)>0$ sufficiently
small so that
$\gamma^r(u)\equiv u+i\gamma_2(u)\in S_{\alpha}$, for any real $u$.
Let us verify
that this path is dissipative for all $\{jk\}\in C_r,\, j<k$.
For $u\in ]-\infty,t_r-L]$, we have, using
\begin{equation}
\mbox{Re }({e}_j(z,0)-{e}_k(z,0))< -2b<0,\;\;\;\forall
z\in S_{\alpha}\cap
\{z|\mbox{Re }z\leq t_r-L\}
\end{equation}
and
\begin{equation}
\left|\mbox{Im }({e}_j(t+is,0)-{e}_k(t+is,0))\right|<
|s|\sup_{s'\in[0,s]}
\left|\mbox{Re }({e}_j'(t+is',0)-{e}_k'(t+is',0))\right|,
\end{equation}
(see (\ref{imre})), and the definition (\ref{defa})
\begin{eqnarray}\label{anco}
& &\mbox{Re }({e}_j(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))\dot{\gamma}_2(u)
+\mbox{Im }({e}_j(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))\dot{\gamma}_1(u)
=\nonumber\\
& &-\mbox{Re }({e}_j(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))\frac{\gamma_2(u)a(u)}{b}+
\mbox{Im }({e}_j(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))>\nonumber\\
& &2\gamma_2(u)a(u)-\gamma_2(u)a(u)/2>\gamma_2(u)a(u)>0.
\end{eqnarray}
Similarly, when $u\geq t_r+L$ we get, using
\begin{equation}
\mbox{Re }({e}_j(z,0)-{e}_k(z,0))> 2b>0,\;\;\;\forall
z\in S_{\alpha}\cap
\{z|\mbox{Re }z\geq t_r+L\},
\end{equation}
\begin{eqnarray}\label{banco}
& &\mbox{Re }({e}_j(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))\dot{\gamma}_2(u)
+\mbox{Im }({e}_1(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))\dot{\gamma}_1(u)
=\nonumber\\
& &\mbox{Re }({e}_j(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))\frac{\gamma_2(u)a(u)}{b}+
\mbox{Im }({e}_j(\gamma^r(u),0)-
{e}_k(\gamma^r(u),0))>\nonumber\\
& &2\gamma_2(u)a(u)-\gamma_2(u)a(u)/2>\gamma_2(u)a(u)>0.
\end{eqnarray}
Finally, for $s\in [t_r-L,t_r+L]$, we have with (\ref{dec})
\begin{eqnarray}\label{bancor}
& &\mbox{Im }({e}_j(\gamma^r(u),0)-{e}_k(\gamma^r(u),0))=
\int_0^{\gamma_2(u)}
\mbox{Re }({e}_j'(t'+is,0)-{e}_k'(t'+is,0))\geq\nonumber\\
& &\gamma_2(u)2c>\gamma_2(u)c>0.
\end{eqnarray}
Thus, $\gamma^r(u)$ is dissipative for all $\{jk\}\in C_r,\, j<k$.
Note that the last estimate shows that it is not possible to find a
dissipative
path for $\{jk\}\in C_r,\, j<k$ below the real axis.
Consider now $\{jk\}\in N,\, j<k$ and let $\gamma_2(u)$ be a solution of
\begin{equation}
\dot{\gamma}_2(u)=-\frac{\gamma_2(u)a(u)}{b}, \;\;\; \gamma_2(0)>0,
\;u\in]-\infty,+\infty [,
\end{equation}
i.e.
\begin{equation}
{\gamma}_2(u)=\gamma_2(0)\mbox{e}^{-\int_{0}^ua(u')du'/b}.
\end{equation}
As above, we have $\gamma_2(u)>0$ for any $u$ and we can choose
$\gamma_2(0)>0$ small enough so that
$\gamma^+(u)\equiv u+i\gamma_2(u)\in S_{\alpha}$ for any $u\in\hbox{$\mit I$\kern-.277em$\mit R$}$.
Since
\begin{equation}
\mbox{Re }({e}_j(z,0)-{e}_k(z,0))> -2b\;\;\; \forall z\in S_{\alpha}
\end{equation}
we check by a computation analogous to (\ref{anco}) that $\gamma^+(u)$
is
dissipative for $\{jk\}\in N,\, j<k$. Similarly, one verifies that if
$\gamma_2(u)$ is the solution of
\begin{equation}
\dot{\gamma}_2(u)=\frac{\gamma_2(u)a(u)}{b}, \;\;\; \gamma_2(0)<0,
\;u\in]-\infty,+\infty [
\end{equation}
with $|\gamma_2(0)|$ small enough, the path $\gamma^-(u)\equiv u+i
\gamma_2(u)$ below
the real axis is in $S_{\alpha}$ for any $u\in\hbox{$\mit I$\kern-.277em$\mit R$}$ and is dissipative for
$\{jk\}\in N,\, j<k$ as well.
Finally, the complex conjugates of these paths yield dissipative paths
above and
below the real axis for $\{jk\}\in N,\, j>k$.
We now define the dissipative domains by means of their
borders. Let $\gamma^+(u)$ and $\gamma^-(u)$, $u\in\hbox{$\mit I$\kern-.277em$\mit R$}$, two dissipative
paths in
$S_{\alpha}$ defined as above with $|\gamma_2^-(0)|$ sufficiently small
so that $\overline{\gamma^-}$ is below $\gamma^+$. We set
\begin{equation}
D=\{z=t+is|0<-\gamma_2^-(t)\leq s\leq\gamma_2^+(t),\,t\in\hbox{$\mit I$\kern-.277em$\mit R$}\}.
\end{equation}
Let $z\in D$, and $j\in \{1,\cdots, n\}$ be fixed. By assumption H5, the
set
$X_j$ of indices $k$ such that
$\{jk\}\in C_r$ for some $r\in \{1,\cdots, p\}$ consists in values $k$
satisfying $j<k$ or
it consists in values $k$ satisfying $j>k$. Let us assume the first
alternative
takes place.
Now for any $k\in \{1,\cdots, n\}$, there are three cases.
\\ 1) If $k\in X_j$, then
there exists a dissipative path $\gamma^r\in D$ for $\{jk\}\in C_r,\,
j<k$
constructed as above
which links $-\infty$ to $z$. It is enough to select the initial condition
$\gamma_2(t_r)$ suitably, see figure \ref{figdis}.
\\ 2) Similarly, if $j<k\not \in X_j$, there
exists a dissipative path $\gamma^+\in D$ for $\{jk\}$ constructed as
above
which links $-\infty$ to $z$ obtained by a suitable choice of
$\gamma_2(0)$.
\\ 3) Finally, if $k>j$, we can take as a dissipative path for $\{jk\}$,
the path
$\overline{\gamma^-}\in D$ constructed as above
which links $-\infty$ to $z$ with suitable choice of $\gamma_2(0)$.
Hence $D$ is
dissipative for the index $j$, when $\delta = 0$.
\begin{figure}
\vspace{5cm}
\hspace{3cm}\special{picture fig6 scaled 700}
\caption{The dissipative domain $D$ and some dissipative paths.}
\label{figdis}
\end{figure}
If $j$ is such that the set $X_j$ consists in points $k$ with $k>j$,
a similar argument
with the complex conjugates of the above paths shows that the domain
$\overline{D}$ below
the real axis is dissipative for $j$ when $\delta = 0$.
Let us show that these domains remain dissipative when $\delta>0$
is not too large. We start by considering the analyticity properties of
the perturbed eigenvalues $e_j(z,\delta)$, $\delta>0$.
Let $0<\alpha'<\alpha$ be such that
\begin{equation}
I_r^{\alpha'}\cap (D\cup \overline{D})=\emptyset\, ,\;\;\;\forall
r=1,\cdots, p.
\end{equation}
The analytic eigenvalues $e_j(z,0)$, $j\in \{1,\cdots ,n\}$, are
isolated in the spectrum of $H(z,0)$ for any
$z\in \widetilde{S}_{\alpha}$,
where
\begin{equation}
\widetilde{S}_{\alpha}={S}_{\alpha}\backslash
\cup_{r=1,\cdots ,p}I_r^{\alpha'}.
\end{equation}
For any $j=1,\cdots, n$ we get from perturbation theory \cite{k}, that
the analytic
continuations $\widetilde{e}_j(z,\delta)$ of $e_j(t_1-L,\delta)$ in
$\widetilde{S}_{\alpha}$ are all distinct in $\widetilde{S}_{\alpha}$,
provided $\delta$ is small enough. This is due to the fact that
assumption H5 implies the continuity of
$H(z,\delta)$ in $\delta$ uniformly in $z\in S_{\alpha}$, as
is easily verified.
More precisely, for any fixed index $j$, the eigenvalue $e_j(t,\delta)$
experiences avoided crossings at the points $\{t_{r_1},\cdots ,t_{r_j}\}$.
We can assume without loss of generality that
\begin{equation}
I_k^{\alpha'}\cap I_l^{\alpha'}=\emptyset \, , \;\;\; \forall k\neq l
\in
\{{r_1},\cdots ,{r_j}\}.
\end{equation}
Hence, for $\delta>0$ small enough, the analytic continuation
$\widetilde{e}_j(z,\delta)$ is isolated in the spectrum of $H(z,\delta)$,
uniformly in $z\in S_{\alpha}\backslash \cup_{r={r_1},\cdots ,{r_j}}
I_r^{\alpha'}$. Since by assumption H5 there is no crossing of
eigenvalues
on the real axis when $\delta>0$, there exists a $0<g<\alpha'$, which
depends on $\delta$,
such that
$\widetilde{e}_j(z,\delta)$ is isolated in the spectrum of $H(z,\delta)$,
uniformly in $z\in {S}^j_{\alpha}$, where
\begin{equation}
{S}^j_{\alpha}={S}_{\alpha}\backslash \cup_{r={r_1},\cdots ,{r_j}}
(J_r\cup\overline{J_r})
\end{equation}
and
\begin{equation}
J_r=I_r^{\alpha'}\cap\{z| \,\mbox{Im }z>g\}\, , \;\;\; r=1,\cdots, p.
\end{equation}
Hence the singularities of $\widetilde{e}_j(z,\delta)$ are located in
$\cup_{r={r_1},\cdots ,{r_j}}(J_r\cup\overline{J_r})$, which yields the
first assertion of the
lemma.
Consider a path $\nu_r$ from $t_r-L$ to $t_r+L$ which goes above $J_r$,
where $t_r$ is an avoided crossing between ${e}_j(t,\delta)$ and
${e}_{k}(t,\delta)$, $k=j\pm 1$.
By perturbation theory again, ${e}_j(t_r-L,\delta)$ and
${e}_k(t_r-L,\delta)$ tend to ${e}_{j'}(t_r-L,0)$ and ${e}_{k'}(t_r-L,0)$
as
$\delta\rightarrow 0$, for some $j',k'\in 1,\cdots, n$, whereas ${e}_j(t_r+L,
\delta)$
and ${e}_k(t_r+L,\delta)$
tend to ${e}_{k'}(t_r+L,0)$ and ${e}_{j'}(t_r+L,0)$ as $\delta\rightarrow 0$,
see figure \ref{figlev}. Now, the analytic continuations
of ${e}_j(t_r-L,\delta)$ and
${e}_k(t_r-L,\delta)$ along $\nu_r$,
$\widetilde{e}_j(z,\delta)$ and
$\widetilde{e}_k(z,\delta)$ tend to the analytic functions
$\widetilde{e}_{j'}(z,0)={e}_{j'}(z,0)$ and
$\widetilde{e}_{k'}(z,0)={e}_{k'}(z,0)$ as
$\delta\rightarrow 0$, for all $z\in\nu_r$. Thus, we deduce that
for $\delta$ small enough
\begin{equation}
\widetilde{e}_j(t_r+L,\delta)\equiv {e}_k(t_r+L,\delta),
\end{equation}
since we know that
$\widetilde{e}_j(t_r+L,\delta)={e}_{\sigma(j)}(t_r+L,\delta)$,
for some permutation $\sigma$.
Hence the point iii) of the lemma follows.
Note that the analytic
continuations $\widetilde{e}_j(z,\delta)$ are single valued in
$\widetilde{S}_{\alpha}$. Indeed, the analytic continuation of
${e}_j(t_r-L,\delta)$ along
$\overline{\nu_r}$, denoted by $\widehat{e}_j(z,\delta)$,
$\forall z\in \overline{\nu_r}$, is such that
\begin{equation}
\widehat{e}_j(t_r+L,\delta)=\overline{\widetilde{e}_j(t_r+L,\delta)}=
\widetilde{e}_j(t_r+L,\delta)={e}_k(t_r+L,\delta),
\end{equation}
due to Schwarz's principle.
We further require $\delta$ to be sufficiently small so that
the following estimates are satisfied
\begin{eqnarray}
& &\label{a}\min_{r\in \{ 1,\cdots,p\}}\min_{\{jk\}\in C_r \atop j<k}
\inf_{z\in
\widetilde{S}_{\alpha}\backslash I_r^{\alpha}}\left|
\mbox{Re }(\widetilde{e}_j(z,\delta)-\widetilde{e}_{k}(z,\delta))
\right|
> b>0\\
& &\label{be}\min_{\{jk\}\in N \atop j<k}\inf_{z\in
\widetilde{S}_{\alpha}}\left|
\mbox{Re }(\widetilde{e}_j(z,\delta)-\widetilde{e}_{k}(z,\delta))
\right|
>b>0\\
& &\label{c}\max_{j<k\in \{ 1,\cdots,n\}}\sup_{\mbox{Im }z |\,z\in
\widetilde{S}_{\alpha}}
\left|\mbox{Re }
(\widetilde{e}_j'(z,\delta)-\widetilde{e}_{k}'(z,\delta))\right|<
a(\mbox{Re }z).
\end{eqnarray}
and, in the compacts $\widetilde{I}_r^{\alpha}={I}_r^{\alpha}\backslash
{I}_r^{\alpha'}$,
\begin{eqnarray}
& &\label{1}
\min_{r\in \{ 1,\cdots,p\}}\min_{\{jk\}\in C_r \atop j<k}
\inf_{z\in\widetilde{I}_r^{\alpha}}|\mbox{Im }
(\widetilde{e}_j(z,\delta)-\widetilde{e}_{k}(z,\delta))|
>\nonumber\\
& &\frac{1}{2}\min_{r\in \{ 1,\cdots,p\}}\min_{\{jk\}\in C_r \atop j<k}
\inf_{z\in\widetilde{I}_r^{\alpha}}|\mbox{Im }
(\widetilde{e}_j(z,0)-\widetilde{e}_{k}(z,0))| > |\mbox{Im }z| c\\
& &\label{2}
\max_{r\in \{ 1,\cdots,p\}}\max_{j<k\in \{ 1,\cdots,n\}}\sup_{z\in
\widetilde{I}_r^{\alpha}}\left|
\mbox{Im }(\widetilde{e}_j(z,\delta)-\widetilde{e}_{k}(z,\delta))
\right|<
\nonumber\\
& &2\max_{r\in \{ 1,\cdots,p\}}\max_{j<k\in \{ 1,\cdots,n\}}
\sup_{z\in
\widetilde{I}_r^{\alpha}}\left|\mbox{Im }
(\widetilde{e}_j(z,0)-\widetilde{e}_{k}(z,0))\right|<
|\mbox{Im }z| a(\mbox{Re }z).
\end{eqnarray}
The simultaneous requirements (\ref{defa}) and (\ref{c}) is made
possible by
the continuity properties of $H'(z,\delta)$ and the uniformity in
$\delta$
of the decay at $\pm\infty$ of $H(z,\delta)$ assumed in H4 together with
the fact that
$a(t)$ can be replaced by a multiple of $a(t)$ if necessary, to satisfy
both estimates.
The condition on $\delta$ is given by the first inequalities in (\ref{1})
and (\ref{2}),
whereas the second ones are just recalls.
Then it remains to check that the paths
$\gamma^r, \gamma^+$ and $\gamma^-$ defined above satisfy the
dissipativity
condition (\ref{disco}) for the corresponding indices.
This is not difficult,
since the above estimates are precisely designed to preserve inequalities
such as
(\ref{anco}), (\ref{banco}) and (\ref{bancor}). However, it should not
be forgotten that in the sets $I_r^{\alpha'}$ the eigenvalues may be
singular so that
(\ref{imre}) cannot be used there. So when checking that a path
parameterized
as above by $u\in \hbox{$\mit I$\kern-.277em$\mit R$}$ is dissipative, it is necessary to consider
separately
the case $u\in \hbox{$\mit I$\kern-.277em$\mit R$}\backslash (\cup_{r=1,\cdots,p}I_r)$, where we proceed
as above with (\ref{a}),
(\ref{be}), (\ref{c}) and (\ref{imre}) and the case
$u\in \cup_{r=1,\cdots,p}I_r$, where we use use (\ref{1}) and (\ref{2})
instead of (\ref{imre}) as follows.
If $u\in I_r$ for $r$ such that $t_r$ is a crossing point for
$e_j(t,0)$ and $e_k(t,0)$, one takes (\ref{1}) to estimate
$\mbox{Im }(\widetilde{e}_{j'}(z,\delta)-\widetilde{e}_{k'}(z,\delta))$
for the corresponding indices $j'$ and $k'$,
and if $t_r$ is not a crossing
point for $e_j(t,0)$ and $e_k(t,0)$, one uses (\ref{2}) to estimate
$\mbox{Im }(\widetilde{e}_{j'}(z,\delta)-\widetilde{e}_{k'}(z,\delta))$.
Consequently, the domains $D$ and
$\overline{D}$ defined above keep the same dissipativity properties when
$\delta>0$ is small enough.
Let us finally turn to the determination of the associated permutation
$\sigma$. As noticed earlier, the eigenvalues $\widetilde{e}_j(z,\delta)$
are continuous in $\delta$, uniformly in $z\in \widetilde{S}_{\alpha}$.
Hence,
since the eigenvalues ${e}_j(z,0)$ are analytic in $S_{\alpha}$, we have
\begin{equation}
\lim_{\delta\rightarrow 0}\widetilde{e}_j(\infty,\delta)={e}_j(\infty,0)
\;\;\; j=1,2,\cdots,n.
\end{equation}
Whereas we have along the real axis (see figure \ref{figlev}),
\begin{equation}
\lim_{\delta\rightarrow 0}{e}_{\sigma(j)}(\infty,\delta)={e}_j(\infty,0),
\end{equation}
with $\sigma$ defined in the lemma,
from which the result follows.\hfill {$\Box$}
\section{Applications}\label{applic}
\setcounter{equation}{0}
Let us consider the time-dependent Schr\"odinger equation in the
adiabatic limit.
The relevant equation is then
(\ref{schr}) where $H(t)=H^*(t)$ is the time-dependent self-adjoint
hamiltonian.
Thus we can take $J=\hbox{$\mit I$\kern-.72em$\mit I$}$ in proposition \ref{PROS} to get
\begin{equation}
H(t)=H^*(t)=H^{\#}(t).
\end{equation}
The norm of an eigenvector being positive, it remains to
impose the gap hypothesis in H3 to fit in the framework and we deduce
that the
$S$-matrix is unitarity, since $R=\hbox{$\mit I$\kern-.72em$\mit I$}$. In this context,
the elements of the $S$-matrix describe the transitions between the
different levels between $t=-\infty$ and $t=+\infty$ in the adiabatic
limit.
We now specify a little more our concern and consider a
three-level system, i.e. $H(t)=H^*(t)\in M_3(\hbox{$\mit I$\kern-.7em$\mit C$})$.
We assume that $H(t)$ satisfies the hypotheses of corollary \ref{PERCO}
with an extra parameter $\delta$ which we omit in the notation and
displays
two avoided crossings at $t_1<t_2$, as shown in figure \ref{adiab}.
\begin{figure}
\vspace{6cm}
\hspace{3cm}\special{picture fig7 scaled 500}
\caption{The pattern of avoided crossings in the adiabatic context.}
\label{adiab}
\end{figure}
The corresponding permutation $\sigma$ is given by
\begin{equation}
\sigma(1)=3,\;\;\; \sigma(2)=1, \;\;\; \sigma(3)=2.
\end{equation}
By corollary \ref{PERCO}, we can compute asymptotically the elements
$s_{31},s_{12},s_{23}$ and
$s_{jj}$, $j=1,2,3$. Using the unitarity of the $S$-matrix, we can get
some more information.
Introducing
\begin{equation}
\Gamma_j=\left|\mbox{Im }\int_{\zeta_j}e_j(z)dz\right| ,\;\;\; j=1,2,
\end{equation}
where $\zeta_j$ is in the upper half plane, with the notation of
section \ref{avoid}, it follows that
\begin{equation}
s_{31}={\cal O} \left(\mbox{e}^{-(\Gamma_1 +\Gamma_2)/\varepsilon}\right), \;\;\;
s_{12}={\cal O}
\left(\mbox{e}^{-\Gamma_1/\varepsilon}\right), \;\;\; s_{23}={\cal O}
\left(\mbox{e}^{-\Gamma_2/\varepsilon}\right)
\end{equation}
and
\begin{equation}
s_{jj}=1+{\cal O} (\varepsilon), \;\;\; j=1,2,3.
\end{equation}
Expressing the fact that the first and second columns as well as the
second and third rows are
orthogonal, we deduce
\begin{eqnarray}
s_{21}&=&-\overline{s_{12}}\frac{s_{11}}{\overline{s_{22}}}
\left(1+{\cal O} \left(\mbox{e}^{-2\Gamma_2/\varepsilon}\right)\right)\\
s_{32}&=&-\overline{s_{23}}\frac{s_{33}}{\overline{s_{22}}}
\left(1+{\cal O} \left(\mbox{e}^{-2\Gamma_1/\varepsilon}\right)\right).
\end{eqnarray}
Finally, the estimate in corollary \ref{PERCO} yields
\begin{equation}
s_{13}={\cal O} \left(\varepsilon\mbox{e}^{-|h|(e_2(\infty,\delta)-e_1(\infty,\delta))/
\varepsilon}\mbox{e}^{-\Gamma_2/\varepsilon}\right)
={\cal O} \left(\mbox{e}^{-(\Gamma_2+\Gamma_2+K)/\varepsilon}\right)
\end{equation}
where $K>0$, since we have that $\Gamma_j\rightarrow 0$ as $\delta\rightarrow 0$.
Hence we get
\begin{equation}
S=\pmatrix{s_{11}&s_{12}&{\cal O} \left(\mbox{e}^{-(\Gamma_2+\Gamma_2+K)/\varepsilon}
\right)\cr
-\overline{s_{12}}\frac{s_{11}}{\overline{s_{22}}}
\left(1+{\cal O} \left(\mbox{e}^{-2\Gamma_2/\varepsilon}\right)\right)
&s_{22}&s_{23}\cr
s_{31}&-\overline{s_{23}}\frac{s_{33}}{\overline{s_{22}}}
\left(1+{\cal O} \left(\mbox{e}^{-2\Gamma_1/\varepsilon}\right)
\right)&s_{33}}
\end{equation}
where all $s_{jk}$ above can be computed asymptotically up to
exponentially small relative error, using
(\ref{imfor}).
The smallest asymptotically computable element $s_{13}$ describes the
transition from $e_1(-\infty,\delta)$ to
$e_3(+\infty,\delta)$. The result we get for this element is in agreement
with the rule
of the thumb claiming
that the transitions take place locally at the avoided crossings and can
be considered as
independent.
Accordingly, we can only estimate the smallest element of all, $s_{13}$,
which
describes the transition from $e_3(-\infty,\delta)$ to
$e_1(+\infty,\delta)$,
for which the avoided crossings are not encountered in "right order",
as discussed in \cite{hp}.
It is possible however to get an asymptotic expression for this element
in some cases.
When the unperturbed levels $e_2(z,0)$ and $e_3(z,0)$ possess a
degeneracy
point in $S_{\alpha}$ and when there exists a dissipative domain for the
index $3$ of the unperturbed
eigenvalues going above
this point, one can convince oneself that $s_{13}$ can be computed
asymptotically
for $\delta$ small enough, using the techniques presented
above.
Our second application is the study of the semi-classical scattering
properties of the multichannel
stationnary Schr\"odinger equation with energy above the potential
barriers.
The relevant equation is then
\begin{equation}\label{muls}
-\varepsilon^2\frac{d^2}{dt^2}\Phi(t)+V(t)\Phi(t)=E\Phi(t),
\end{equation}
where $t\in\hbox{$\mit I$\kern-.277em$\mit R$}$ has the meaning of a space variable, $\Phi(t)\in\hbox{$\mit I$\kern-.7em$\mit C$}^m$
is
the wave function, $\varepsilon\rightarrow 0$ denotes
Planck's constant, $V(t)=V^*(t)\in M_m(\hbox{$\mit I$\kern-.7em$\mit C$})$ is the matrix of potentials
and
the spectral parameter $E$ is kept fixed and large enough so that
\begin{equation}\label{cosp}
U(t)\equiv E-V(t)>0.
\end{equation}
Introducing
\begin{equation}
\psi(t)=\pmatrix{\Phi(t)\cr i\varepsilon \Phi(t)}\in \hbox{$\mit I$\kern-.7em$\mit C$}^{2m},
\end{equation}
we cast equation (\ref{muls}) into the equivalent form (\ref{schr})
for $\psi(t)$ with generator
\begin{equation}\label{gess}
H(t)=\pmatrix{{\bf O} & \hbox{$\mit I$\kern-.72em$\mit I$}\cr
U(t)& {\bf O}}\in M_{2m}(\hbox{$\mit I$\kern-.7em$\mit C$}).
\end{equation}
It is readily verified that
\begin{equation}
H(t)=J^{-1}H^*(t)J
\end{equation}
with
\begin{equation}
J=\pmatrix{{\bf O} & \hbox{$\mit I$\kern-.72em$\mit I$}\cr
\hbox{$\mit I$\kern-.72em$\mit I$} & {\bf O}}.
\end{equation}
Concerning the spectrum of $H(t)$, it should be remarked that if the
real and positive eigenvalues of $U(t)$, $k_j^2(t)$, $j=1,\cdots, m$
associated with
the eigenvectors $u_j(t)\in\hbox{$\mit I$\kern-.7em$\mit C$}^m$ are
assumed to be distinct, i.e.
\begin{equation}
0<k_1^2(t)<k_2^2(t)<\cdots <k_m^2(t),
\end{equation}
then the spectrum of the generator $H(t)$ given by (\ref{gess}) consists
in
$2m$ real distinct eigenvalues
\begin{equation}\label{not1}
-k_m(t)<-k_{m-1}(t)<\cdots <-k_1(t)<k_1(t)<k_2(t)<\cdots <k_m(t)
\end{equation}
associated with the $2m$ eigenvectors
\begin{eqnarray}\label{not2}
& &\chi_j^{\pm}(t)=\pmatrix{u_j(t)\cr\pm k_j(t)u_j(t)}\in\hbox{$\mit I$\kern-.7em$\mit C$}^{2m},
\nonumber\\
& &H(t)\chi_j^{\pm}(t)=\pm k_j(t)\chi_j^{\pm}(t).
\end{eqnarray}
We check that
\begin{equation}\label{nvp}
(\chi_j^{\pm}(0),\chi_j^{\pm}(0))_J=\pm2k_j(0)\|u_j(0)\|\neq 0,\;\;\;
j=1,\dots ,m\,
\end{equation}
where $\|u_j(0)\|$ is computed in $\hbox{$\mit I$\kern-.7em$\mit C$}^m$, so that proposition \ref{PROS}
applies. Before dealing with its consequences,
we further explicit the structure of $S$.
Adopting the notation suggested by (\ref{not1}) and (\ref{not2}) we
write
\begin{eqnarray}
H(t)&=&\sum_{j=1}^nk_j(t)P_j^+(t)-\sum_{j=1}^nk_j(t)P_j^-(t)\\
\label{decpm}
\psi(t)&=&\sum_{j=1}^nc^+_j(t)\varphi_j^+(t)\mbox{e}^{-i\int_0^tk_j(t')dt'/\varepsilon}
+
\sum_{j=1}^nc^-_j(t)\varphi_j^-(t)\mbox{e}^{i\int_0^tk_j(t')dt'/\varepsilon}
\end{eqnarray}
and introduce
\begin{equation}
{\bf c}^{\pm}(t)=\pmatrix{c_1^{\pm}(t)\cr c_2^{\pm}(t)\cr \vdots \cr
c_m^{\pm}(t)}\in \hbox{$\mit I$\kern-.7em$\mit C$}^m.
\end{equation}
Hence we have the block structure
\begin{equation}\label{sm}
S\pmatrix{{\bf c}^{+}(-\infty)\cr {\bf c}^{-}(-\infty)}\equiv
\pmatrix{S_{++}&S_{+-}\cr S_{-+}&S_{--}}\pmatrix{{\bf c}^{+}(-\infty)
\cr {\bf c}^{-}(-\infty)}=
\pmatrix{{\bf c}^{+}(+\infty)\cr {\bf c}^{-}(+\infty)}
\end{equation}
where $S_{\sigma \tau}\in M_{m}(\hbox{$\mit I$\kern-.7em$\mit C$})$, $\sigma, \tau\in\{+,-\}$.
Let us turn to the symmetry properties of $S$. We get from (\ref{nvp})
and
proposition \ref{PROS} that
\begin{equation}
\pmatrix{S_{++}&S_{+-}\cr S_{-+}&S_{--}}^{-1}=
\pmatrix{\hbox{$\mit I$\kern-.72em$\mit I$} & {\bf O} \cr {\bf O} & -\hbox{$\mit I$\kern-.72em$\mit I$}}
\pmatrix{S_{++}&S_{+-}\cr S_{-+}&S_{--}}^*
\pmatrix{\hbox{$\mit I$\kern-.72em$\mit I$} & {\bf O} \cr {\bf O} & -\hbox{$\mit I$\kern-.72em$\mit I$}}=
\pmatrix{S_{++}^*&-S_{-+}^*\cr -S_{+-}^*&
S_{--}^*}.
\end{equation}
In terms of the blocks $S_{\sigma \tau}$, this is equivalent to
\begin{eqnarray}\label{222}
& &S_{++}S_{++}^*-S_{+-}S_{+-}^*=\hbox{$\mit I$\kern-.72em$\mit I$}\\ \label{223}
& &S_{++}S_{-+}^*-S_{+-}S_{--}^*={\bf O}\\
& &S_{--}S_{--}^*-S_{-+}S_{-+}^*=\hbox{$\mit I$\kern-.72em$\mit I$} .
\end{eqnarray}
The block $S_{++}$ describes the transmission coefficients associated
with a wave traveling
from the right and $S_{-+}$ describes the associated reflexion
coefficients.
Similarly, $S_{--}$ and $S_{+-}$ are related to the transmission and
reflexion
coefficients
associated with a wave incoming from the left. It should be noted that in
case of
equation (\ref{muls}) another convention is often used to define an
$S$-matrix,
see \cite{f1}, for instance. This gives rise to a different $S$-matrix
with
similar interpretation. However it is not difficult to establish a
one-to-one correspondence between the two definitions.
If the matrix of potentials $V(t)$ is real symmetric, we have further
symmetry
in the $S$-matrix.
\begin{lem}\label{REPOT}
Let $S$ given by (\ref{sm}) be the $S$-matrix associated with
(\ref{muls}) under condition (\ref{cosp}). If we further assume
$V(t)=\overline{V(t)}$, then, taking $\varphi_j^{\pm}(0)\in \hbox{$\mit I$\kern-.277em$\mit R$}^{2m}$,
$j=1,\cdots, m$, in
(\ref{decpm}), we get
\begin{equation}
S_{++}=\overline{S_{--}}\, ,\;\;\; S_{+-}=\overline{S_{-+}}.
\end{equation}
\end{lem}
The corresponding results for the $S$-matrix defined in \cite{f1} are
derived
in \cite{mn}. The proof of this lemma can be found in appendix.
We consider now (\ref{muls}) the case
$U(t)=U^*(t)=\overline{U(t)}\in M_2(\hbox{$\mit I$\kern-.277em$\mit R$})$, which
describes a two-channel Schr\"odinger equation. We assume that the
four-level generator
$H(t)$ displays three avoided crossings at $t_1<t_2$, two of which
take place at the same
point $t_1$, because of the symmetry of the eigenvalues, as in
figure \ref{semic}.
\begin{figure}
\vspace{7cm}
\hspace{3cm}\special{picture fig8 scaled 500}
\caption{The pattern of avoided crossings in the semiclassical context.}
\label{semic}
\end{figure}
By lemma \ref{REPOT}, it is enough to consider the blocks $S_{++}$ and
$S_{+-}$.
The transitions corresponding to elements of these blocks which we can
compute
asymptotically are from level $1^+$ to level $2^+$ and from level $2^-$
to level $1^+$.
They correspond to elements $s^{++}_{21}$ and $s^{+-}_{12}$ respectively.
With the
notations
\begin{equation}
\Gamma_j=\left|\mbox{Im }\int_{\zeta_j}k_1(z)dz\right| ,\;\;\; j=1,2,
\end{equation}
where $\zeta_j$ is in the upper half plane,
we have the estimates
\begin{equation}\label{esss}
s^{++}_{21}={\cal O} \left(\mbox{e}^{-\Gamma_1/\varepsilon}\right), \;\;\;
s^{+-}_{12}={\cal O} \left(\mbox{e}^{-(\Gamma_1 +\Gamma_2)/\varepsilon}\right),\;\;\;
s^{++}_{jj}=1+{\cal O} (\varepsilon), \;\;\; j=1,2.
\end{equation}
It follows from (\ref{223}) and lemma \ref{REPOT} that the matrix
$S_{++}S_{+-}^T$
is symmetric. Hence
\begin{equation}\label{resly}
s^{++}_{11}s^{+-}_{21}+s^{++}_{12}s^{+-}_{22}=
s^{++}_{21}s^{+-}_{11}+s^{++}_{22}s^{+-}_{12},
\end{equation}
whereas we get from (\ref{222})
\begin{equation}\label{from}
s^{++}_{11}\overline{s^{++}_{21}}+s^{++}_{12}\overline{s^{++}_{22}}=
s^{+-}_{11}\overline{s^{+-}_{21}}+s^{+-}_{12}\overline{s^{+-}_{22}}.
\end{equation}
The only useful estimate we get with corollary \ref{PERCO} is
\begin{equation}
s^{+-}_{22}={\cal O} \left(\mbox{e}^{-(\Gamma_1 +\Gamma_2+K)/\varepsilon}\right),\;\;\;
K>0,
\end{equation}
which yields together with (\ref{esss}) in (\ref{resly})
\begin{equation}
s^{+-}_{21}=s^{++}_{21}s^{+-}_{11}/s^{++}_{11}+{\cal O}
\left(\mbox{e}^{-(\Gamma_1 +\Gamma_2)/\varepsilon}\right).
\end{equation}
Thus, from (\ref{from}) and (\ref{sjksup}) for $s^{+-}_{11}$,
\begin{equation}
s^{++}_{12}=-\overline{s^{++}_{21}}\frac{s^{++}_{11}}
{\overline{s^{++}_{22}}}
\left(1+{\cal O} \left(\mbox{e}^{-\kappa/\varepsilon}\right)\right),
\end{equation}
with
\begin{equation}\label{bbeh}
0<\kappa<\min(\Gamma_1,\Gamma_2).
\end{equation}
Summarizing, we have
\begin{equation}
S_{++}=\pmatrix{s^{++}_{11}&-\overline{s^{++}_{21}}\frac{s^{++}_{11}}
{\overline{s^{++}_{22}}}
\left(1+{\cal O} \left(\mbox{e}^{-\kappa/\varepsilon}\right)\right)\cr
s^{++}_{21}&s^{++}_{22}}
\end{equation}
and
\begin{equation}
S_{+-}=\pmatrix{{\cal O} \left(\mbox{e}^{-\kappa/\varepsilon}\right) &s^{+-}_{12}\cr
{\cal O} \left(\mbox{e}^{-\kappa/\varepsilon}\right) &{\cal O} \left(
\mbox{e}^{-(\Gamma_1 +\Gamma_2+K)/\varepsilon}\right)},
\end{equation}
where all elements $s^{\sigma\tau}_{jk}$ can be asymptotically computed
up to exponentially
small relative corrections using (\ref{imfor}). We get no information on
the first
column of $ S_{+-}$ but the estimate (\ref{sjksup}) where necessarily,
(\ref{bbeh}) holds.
However, if there exists one or several
other dissipative domains for certain indices, it is then possible to get
asymptotic formulas for
the estimated terms.
|
\section{Six-Quark Models for the d'-Dibaryon}
\noindent
In the bag-string model, dibaryons are described as rotating
strings with colored quark clusters at the ends \cite{Mul78}.
This model leads to a linear Regge trajectory of excited
states. It predicts that the lowest $L=1$ excited state
is obtained for a diquark-tetraquark configuration at around 2100 MeV
\cite{Mul78}.
\begin{figure}[htb]
\label{Fig.1}
$$\mbox{
\epsfxsize 10.0 true cm
\epsfysize 5.5 true cm
\setbox0= \vbox{
\hbox { \centerline{
\epsfbox{fig1ep.ps}
} }
}
\box0
} $$
\vspace{0.1cm}
\caption[The d'-dibaryon in the quark cluster model]
{The d'-dibaryon in the $q^4$-$q^2$ quark cluster model.}
\end{figure}
The drawback of this approach is that it does not respect
the Pauli principle. Only the quarks within
the individual clusters are antisymmetrized but not the quarks
belonging to different clusters. This is a good approximation
for high angular momentum states because in this case
the system is fairly elongated and the probability of
cluster overlap is small. On the other hand, for a low lying
$L=1$ excitation, such as the $d'$ one expects a considerable amount
of quark exchange between the two clusters.
\begin{table}[htb]
\caption{The mass ($M_{d'}$) and size ($b_6$) of the $d'$
in the quark cluster model. The masses of the diquark and tetraquark
are also given.}
\begin{center}
\begin{tabular}{|c||c|c|c|c|}\hline
& $M_{2q}$ & $M_{4q}$ & $M_{d'}$ & $b_6$\\
Set I & 645 & 1455 & 2440& 0.76 \\
Set II & 637 & 1501 & 2634 & 0.70 \\
Set III & 621 & 1309 & 2111 & 0.95 \\
\hline
\end{tabular}
\end{center}
\end{table}
The confining forces between the colored quarks
prevent large separations of the clusters and the typical size of
such a system is expected to be about 1 fm.
{}From our experience with the $NN$ system we know that the Pauli
principle plays an important role at such distances \cite{Yam91}.
\vspace{0.5cm}
\noindent
{\it 3.1~ The Quark Cluster Model of the $d'$-Dibaryon }
\vspace{0.5cm}
\nobreak
\noindent
In this model, the $d'$ is described as
a nonrelativistic six-quark system in which the quarks
interact via the two-body potentials
of eq.(\ref{Ham}). Tensor and spin-orbit interactions have been omitted
since it has previously been shown
that they give a negligible
contribution to the $d'$ mass \cite{Glo94,Wag95}.
The six-quark wave function is expanded into the cluster
basis
\begin{eqnarray}
\label{rgmwf}
\mid \Psi_{d'}^{J=0,T=0}>
& = & {\cal A} {\Bigl \vert}
\Biggl [ \Bigl [ \Phi_{T}^{S_T=1,T_T=0}
(\b{\rho}_{T},\b{\lambda}_{T},\b{\eta}_T)
\times \ \Y{211}_C
\nonumber \\
& & \otimes\Phi_{D}^{S_D=0,T_D=0} (\b{\rho}_{D}) \times \Y{11}_C
\Bigr ]^{S=1,T=0}
\otimes\chi_{L=1}({\bf R}) \Biggr ]^{J=0,T=0} \ \Y{222}_C \Bigr >,
\end{eqnarray}
where
$\Phi_T^{S_T=1,T_T=0}(\b{\rho}_{T},\b{\lambda}_{T},\b{\eta}_T)$
and
$\Phi_B^{S_D=0,T_D=0}(\b{\rho}_{D})$
are the internal wave functions of the tetraquark (T)
and diquark (D) clusters, respectively and
$\chi_{L=1}({\bf R})$ is the wave function of the relative motion of the
two clusters.
We use the same harmonic oscillator parameter for
the internal and relative motion wave functions.
The Young diagrams in eq.(\ref{rgmwf}) show
that two color triplet clusters $[211]_C$ and $[11]_C$
are coupled to a $[222]_C$ color-singlet six-quark state.
Furthermore, they show that the tetraquark and $d'$ wave function
are not fully antisymmetric but have mixed symmetry in color space.
This nonfactorizability of the color space
considerably complicates the calculation.
\begin{figure}[htb]
\label{Fig.2}
$$\mbox{
\epsfxsize 11.5 true cm
\epsfysize 6.5 true cm
\setbox0= \vbox{
\hbox { \centerline{
\epsfbox{fig2ep.ps}
} }
}
\box0
} $$
\vspace{0.2cm}
\caption[Potential matrix elements]
{The direct, one-quark and two-quark
exchange diagrams that have to be evaluated for each two-body potential.
The horizontal bars indicate the confinement, the
one-gluon, one-pion, or one-sigma
exchange interactions in eq.( \ref{Ham}).}
\end{figure}
\par
The advantage of the cluster model is that it provides
a continuous transition from the $q^6$ six-quark state to the
$q^4-q^2$ clusterized state by smoothly going through all intermediate
configurations. There is no rigid and artificial boundary between
these extreme configurations;
everything is contained in one and the same RGM wave function.
This important property is a direct consequence of the Pauli principle
on the quark level, which is ensured by the
antisymmetrizer ${\cal A}$
\begin{equation}
\label{ant}
{\cal A}= 1 - 8P_{46}^{OSTC} +6P_{35}^{OSTC} P_{46}^{OSTC},
\end{equation}
where $P_{ij}^{OSTC}$ is the permutation operator of
the i-th and j-th quark in orbital (O)
spin-isospin (ST) and color space (C), respectively.
The direct, as well as the one- and
two-quark exchange contributions for the two-body potential of
eq.(\ref{Ham}) are depicted in fig.2.
The solution for the unknown relative wave function $\chi_L({\bf R})$
and the unknown eigenenergy is obtained from the variational principle
\begin{equation}
\delta\left[{ \langle\Psi_{d'}\vert H-E\vert \Psi_{d'} \rangle
\over\langle\Psi_{d'}\vert \Psi_{d'}\rangle }\right ]=0,
\end{equation}
where the variation is with respect to the relative wave function
$\chi_L ({\bf R}) $. The results for the energy (mass) of the $d'$
as well as for the harmonic oscillator parameter
$b_6$ which minimizes the $d'$ mass are shown in table 2 for the parameter
sets of table 1.
\vspace{0.5cm}
\noindent
{\it 3.2~ Shell-Model Calculation for a J$^P$=0$^-$, T=0 six-quark system}
\vspace{0.5cm}
\noindent
Next, we calculate the mass of the
$d'$-dibaryon in the translationally invariant
shell-model (TISM) \cite{Glo94,Wag95}.
Due to the negative parity of the $d'$, only an odd
number of oscillator quanta
$N=1,3,5,...$ is allowed.
There is only one $N=1$ state
which is compatible with $J^p=0^-,T=0$
\begin{equation}
\label{smgs}
\mid \Psi_{d'_{g.s.}}> =
\mid N=1, [51]_O, (\lambda\mu)=(10), L=1, S=1, T=0, [321]_{ST}>.
\end{equation}
For an unambigious classification of TISM states one has
to specify the number of internal
excitation quanta $N$, the Elliot symbol $(\lambda\mu)$, the Young
pattern $[f]_O$ of the spatial permutational
$S_6$-symmetry, further the total orbital angular momentum $L$,
total spin $S$ and total isospin $T$ of the system.
The specification of the intermediate $SU(4)_{ST}$ symmetry is
necessary because in general, the same symmetry in $STC$ space can be
obtained from several states with different intermediate $ST$ symmetries.
The mass of the $d'$ is then given in first order perturbation theory
by the expectation value of the Hamiltonian between the
lowest harmonic oscillator state
of eq.({\ref{smgs})
\begin{equation}
\label{smm}
M_{d'}(b_6) = <\Psi_{d'_{g.s.}} \mid H \mid \Psi_{d'_{g.s.}}>.
\end{equation}
In order to estimate the effect of configuration mixing
with excited shell model states we include
in addition ten $N=3$ states with orbital $[42]_O$ symmetry \cite{Wag95}.
In this case also the $[51]_{ST}$, $[411]_{ST}$, $[33]_{ST}$,
$[321]_{ST}$, and $[2211]_{ST}$ $S_6$ permutational symmetries are allowed.
\par
With fixed parameters of the quark-quark interaction
determined from eq.(\ref{constraints}) we minimize
the $d'$ mass with respect to the harmonic oscillator parameter $b_6$
in the six-quark wave function. Note, that the harmonic oscillator
parameter of the single baryon ($b$) and
the $d'$ ($b_6$) wave function are different.
The value of $b_6$ which minimizes the $d'$ mass
is a measure of the size of the system and is
also given in table 3.
\begin{table}[htb]
\caption{The mass ($M_{d'}$) and size ($b_6$) of the $d'$
in the six-quark shell model without and with configuration mixing.}
\begin{center}
\begin{tabular}{|c||c|c||c|c|}\hline
& \multicolumn{2}{c||}{$N=1$} & \multicolumn{2}{c|}{$N=1$ $\&$ $N=3$ } \\
\hline
& $M^{(N=1)}_{d'}$ [MeV] & $b_6$ [fm] & $M_{d'}$ [MeV] & $b_6$ [fm] \\
\hline
Set I & 2484 & 0.78 & 2413 & 0.78 \\
Set II & 2636 & 0.72 & 2553 & 0.73 \\
Set III & 2112 & 0.95 & 2063 & 0.96 \\
\hline
\end{tabular}
\end{center}
\end{table}
\section{ Discussion and Summary}
\noindent
As is evident from tables 2 and 3 the $d'$ mass of the cluster model
is lower than the single $N=1$ shell-model mass of eq.(\ref{smm})
but higher than the shell-model result with
configuration mixing. Note that the $d'$ mass calculated with set II
(without pion and sigma-exchange between quarks)
is some 150-200 MeV higher than the result with chiral interactions (set I).
In any case, the calculated mass is about 350 MeV higher than the
value required by experiment. However, the confinement strength $a_c$
in the three-quark and six-quark system need not be the same.
If we assume (set III) that $a_c$ in the six-quark system is weaker
than in the nucleon one obtains considerably
smaller values for $M_{d'}$. This assumption is supported by the
harmonic oscillator relation for $a_c$
\begin{equation}
\label{confstr}
a_c \propto {1\over m_q b^4} {1\over N}
\end{equation}
which is inversely proportional to the number of quarks $N$ in the system.
A weaker confinement strength is also expected due to the larger
hadronic size of the $d'$ ($b_6$) as compared to the
hadronic size of the nucleon $(b)$.
Set III differs from Set I of table 1 only in the strength of
the parameter $a_c$ for which we take the value
$a_c=5.0$ MeV/fm$^2$ in the six-quark calculation.
Finally, both calculations
give similar results for the
$d'$ mass and for its size.
Let us briefly discuss the reasons for this.
The outer product of the orbital
$[4]_O$ (tetraquark) and $[2]_O$ (diquark)
symmetries gives the following six-quark symmetries
\begin{equation}
[4]_O \otimes [2]_O = [42]_O \ \ \oplus \ \ [51]_O \ \ \oplus \ \ [6]_O.
\end{equation}
With the exception of the $[6]_O$ symmetry which is incompatible with
$d'$ quantum numbers these are
also included in the enlarged $N=3$ shell-model basis \cite{Wag95}.
Analogously, the outer product of the two clusters in spin-isospin space
leads to
\begin{equation}
\label{STSYM}
[31]_{ST} \otimes [2]_{ST} = [51]_{ST} \ \ \oplus \ \ [42]_{ST} \ \
\oplus \ \ [33]_{ST} \ \ \oplus \ \ [411]_{ST} \ \ \oplus \ \ [321]_{ST}.
\end{equation}
Comparison with eq.(10) in ref.\cite{Glo94} shows
that the $q^4-q^2$ cluster model wave function comprises
the same $S_6$-symmetries in orbital and spin-isospin space
(with the exception of the $[2211]_{ST}$ symmetry) as our
enlarged shell model basis. Thus the trial function space spanned by
both sets of basis functions is not very different.
\par
In summary, we have calculated the mass of a $J^P=0^-$ $T=0$
six-quark system in the NRQM using two different assumptions
for the spatial distribution of the six quarks. The parameters
have been determined from the constraints of eq.(\ref{constraints}).
As in our previous works \cite{Glo94,Wag95}
our results are typically 300-400 MeV above the
required resonance energy. However, for a weaker
confinement strength $a_c$ in the six-quark system as suggested by
eq.(\ref{confstr}) we find a mass for the $d'$ that is considerably smaller.
The assumption of a weaker confinement strength in the six-quark system
does not affect previous results of the model in the $B=2$ sector such as
$NN$ scattering phase shifts or deuteron electromagnetic form factors
which are completely insensitive to the model and strength
of confinement \cite{Shi89}.
|
\section*{Acknowledgements}
This work grew out of a collaboration with Maarten Bergvelt \cite{berg}, and
I have enjoyed discussions with him and with Mitchell Rothstein.
|
\part{ Z(\theta)=\int \prod_n dz_n d{\overline z_n} e^{-S_\theta}/
\int \prod_n dz_n d{\overline z_n} e^{-S}
}
and the free energy density $F(\theta)$ is given by
\eqn\fre{
F(\theta)=-{1 \over V }\log Z(\theta),
}
where $V$ is volume of the system.
The topological charge $\hat{Q}$ is the number of times the fields
cover the sphere $S^2$. The lattice counterpart we adopt is that of the
geometrical definition in ref. \ref\BL
{B.~Berg and M.~L\"uscher, \NP{190[FS3]}, 412 (1981). };
the charge density $\hat{Q}(n^*)$ at dual site $n^*$ is given by
\eqnn\tpc$$\eqalignno{
\hat{Q}(n^*)={1 \over 2\pi }{\rm Im} \Bigl\{ &\ln
\big[{\rm Tr} P(n)P(n+1)(P(n+1+2)\big] \cr &+
\ln \big[{\rm Tr }P(n)P(n+1+2)P(n+2)\big]\Bigr\},&\tpc\cr}$$
where $P(n)_{\alpha\beta} = z_{\alpha n} {\overline z_{\beta n}}$ and
$n$ is the left corner of the plaquette with center $n^*$.
This amounts, in terms of $z$, to the topological charge
\eqn\tpcc{
\hat{Q}={1 \over 4\pi } \sum_{n,\mu,\nu} \epsilon_{\mu \nu}
( \theta_{n,\mu} + \theta_{n+\mu,\nu} -
\theta_{n+\nu,\mu} - \theta_{n,\nu} ),
}
where $\theta_{n,\mu}= {\rm arg} \{ {\overline z_n}z_{n+\mu} \}$.\par
In order to simulate the model with the complex Boltzmann factor,
we follow the Wiese's idea \W.
It, in principle, introduces the constrained updating
of the fields, in which the topological charge,
being a functional of the dynamical fields, is constrained to take a given
value $Q$.
So the phase factor $e^{i \theta Q}$ is factored out, so that
the partition function is given by the summation
of the probability distribution $P(Q)$ weighted by $e^{i \theta Q}$
in each $Q$ sector.
This amounts, in practice, to calculate first the probability
distribution $P(Q)$ at $\theta=0$ and to be followed by taking the
Fourier transform of $P(Q)$ to get the partition function $Z(\theta)$ as
\eqn\four{
Z(\theta)=\sum_Q P(Q) e^{ i \theta Q},
}
where $P(Q)$ is
\eqn\tcd{
P(Q)={\int \prod_n dz_n d{\overline z_n} ^{(Q)} e^{-S}
\over \int \prod_n dz_n d{\overline z_n} e^{-S} }.
}
Here $\prod_n dz_n d{\overline z_n} ^{(Q)} = \prod_n dz_n
d{\overline z_n} \delta_{\hat{Q},Q}$, i.e. , the integration measure
restricted to the configurations with given $Q$, where $\delta_{\hat{Q},Q}$
is the
Kronecker's delta. Note that $\sum_Q P(Q)=1$.
Expectation value of an observable $O$ is given in terms of $P(Q)$ as
\eqn\evq{
\langle O \rangle _{\theta} = {\sum_Q P(Q) \langle O \rangle_Q
e^{ i \theta Q}
\over
\sum_Q P(Q) e^{i \theta Q} },
}
where $\langle O \rangle_Q$ is the expectation value of $O$ at $\theta=0$
for a given $Q$ sector
\eqn\epv{
\langle O \rangle _Q ={\int \prod_n dz_n d{\overline z_n}^{(Q)} O e^{-S}
\over
\int \prod_n dz_n d{\overline z_n}^{(Q)} e^{-S} }.
}
\smallskip
\subsec{algorithm}
We measure the topological charge distribution $P(Q)$ by Monte Carlo
simulation by the Boltzmann weight $\exp(-S)$, where $S$ is defined
by \act.
The standard Metropolis method is used to
update configurations. To calculate $P(Q)$ effectively, we apply
(i) the set method and (ii) the trial distribution method simultaneously.
In the following, we explain briefly the algorithm to make the paper
self-contained.
All we have to calculate $P(Q)$ is to count how many
times the configuration of $Q$ is visited by the histogram method.
The distribution $P(Q)$ could damp very rapidly as $\vert Q \vert$ becomes
large.
We need to calculate the $P(Q)$ at large $\vert Q \vert$'s which would
contribute to $F(\theta)$, $\langle Q \rangle_\theta$ and
$\langle Q^2 \rangle_\theta$ because
they are obtained by the Fourier transformation of $P(Q)$
and its derivatives.
Further, the error of $P(Q)$ at large
$\vert Q \vert$ must be suppressed as small as possible. These are reasons
why we apply two techniques mentioned above.
Since $P(Q)$ is analytically shown to be even function and is certified by
simulation, we restrict to the range of $Q$ to $\geq 0$.\par
The range of $Q$ is grouped into sets $S_i$ ; $S_1 (Q=0 \sim 3)$,
$S_2 (Q=3 \sim 6)$, $\cdots$ , $S_i (Q=3(i-1) \sim 3i)$, $\cdots$ (set
method).
Monte Carlo updatings are done as follows by starting from a
configuration within a fixed set $S_i$.
When $Q$ of a trial configuration ${C_t}$
stays in one of the bins within $S_i$, the configuration ${C_t}$
is accepted, and the count of the corresponding $Q$ value is increased by one,
while when ${C_t}$ goes out of the set $S_i$, ${C_t}$ is rejected,
and the count of $Q$ value of the old configuration is increased by one.
This is done for all sets $S_i$ ; $i$ = 1, 2, $\cdots$.\par
Another of the two techniques is to modify the Boltzmann weight by introducing
trial distributions $P_t(Q)$ for each set (trial distribution method).
This is to remedy $P(Q)$ which falls too rapidly even within a set
in some cases.
We make the counts at
$Q=3(i-1), 3(i-1)+1, 3(i-1)+2$ and $3i$ in each set $S_i$ almost the same.
As the trial distributions $P_t(Q)$'s, we apply the form
$$
P_t(Q) = A_i \exp [ - \bigl( C_i(\beta) / V \bigr) Q^2 ],
$$
where the value of $C_i(\beta)$ depends on the set $S_i$,
and $A_i$ is a constant.
That is, the action during updatings is modified to the effective one
such as $S_{\rm eff}=S + \log P_t(Q) $.
We adjust
$C_i(\beta)$ from short runs to get almost flat distribution
at every $Q$ in $S_i$. \par
To obtain the normalized distribution $P(Q)$ in the whole range of $Q$
from the counts at each set, we make matchings as follows:
\itemitem{i).} At each set $S_i$ ( $i$ = 1, 2, $\cdots$ ), the number of counts
is multiplied
by $P_t(Q)$ at each $Q$. We call the multiplied value $N_i(Q)$,
which is hopefully proportional to the desired topological charge distribution
$P(Q)$.
\itemitem{ii).} In order to match the values in two neighboring sets $S_i$ and
$S_{i+1}$,
we rescale $N_{i+1}(Q)$ so that $N_{i+1}(Q) \rightarrow N_{i+1}(Q) \times r$,
where $r$ =$N_i(Q=3i) / N_{i+1}(Q=3i)$, the ratio of the number of counts at
the right edge of $S_i$ to that at the left edge of $S_{i+1}$.
These manipulations are performed over all the sets.
\itemitem{iii).}The rescaled $N_i(Q)$'s are normalized to
obtain $P(Q)$ such that
$$
P(Q) = { N_i(Q) \over \sum_i\sum_Q N_i(Q) }.
$$
\bigskip
\newsec{Numerical Results}
We use square lattices with the periodic boundary conditions.
Lattice sizes are $V = L \times L$, and $L$ ranges from $L =$ 24, 36, 48 to
72. The total number of counts in each set is $10^4$. The error analysis is
discussed in Appendix.
To check the algorithm, we calculated the internal energy. It agrees with
the analytical results of the strong and weak coupling expansions \BL.
Using the calculated $P(Q)$, we will estimate the free energy $F(\theta)$
and its derivative $\langle Q \rangle_\theta$, respectively.
\smallskip
\subsec{topological charge distribution $P(Q)$}
In this subsection we discuss the topological charge distribution $P(Q)$.
Partition function can be given by the measured $P(Q)$ as in $\four$
in principle, but
we should be careful for estimating $Z(\theta)$ from $P(Q)$.
Since $P(Q)$ is very sharply decreasing function of $Q$, its Fourier series
$Z(\theta)$ is
drastically affected by statistical fluctuations of $P(Q)$.
For example, consider two different $Q$ values, say, $Q_1$ and $Q_2$
($Q_1 \ll Q_2$).
Small error $\delta P(Q_1)$ at $Q_1$ could cause very large effects
to $Z(\theta)$ because $P(Q_2)$ itself is sometimes much smaller than $\delta
P(Q_1)$.
So the effort to obtain $P(Q)$ at large $Q$ may be useless if we allow these
fluctuations
at the small value of $Q$. In order to avoid this problem, we first fit the
measured $P(Q)$ by the
appropriate functions $P_{\rm fit}(Q)$ and obtain $Z(\theta)$ using Fourier
transforming
from $P_{\rm fit}(Q)$. We apply the chi-square-fitting
to the logarithm of the measured $P(Q)$ in the form of
polynomial functions of $Q$
$$
P(Q) = \exp \Big[ \sum_n a_n Q^n \Big].
$$
In the following, we present the results of $\beta$ and volume dependence of
$P(Q)$ .
\par
In Fig.1, we show the measured $P(Q)$ for various $\beta$'s ($\beta =$ 0.0,
0.5, $\cdots$, 3.5)
for a fixed volume ($L=24$). As $\beta$ varies, $P(Q)$ smoothly changes from
strong to weak coupling regions. In the strong coupling regions $(\beta \lsim
2.0)$,
$P(Q)$ shows Gaussian behavior. In the weak coupling regions
$(2.75 \lsim \beta)$,
$P(Q)$ deviates gradually from the Gaussian form, being enhanced at large $Q$
compared to the Gaussian. In order to investigate the difference between the
two regions
in detail, we use the chi-square-fitting to $\log P(Q)$. Table I shows the
results of the fittings, i.e.,
the coefficients $a_n$ of the used polynomial $\sum_n a_n Q^n$ for various
$\beta$'s with
the resulting $\chi^2/d.o.f$. (i) For $\beta \lsim 2.0$, $P(Q)$'s are
indeed fitted well by the Gaussian
form. (ii) For $\beta \gsim 2.75$, terms up to quartic one
are needed for sufficiently good fitting. The linear term, in particular,
is important for
fitting the data at very small $Q$ values.
The value $Q_{\rm Max}$, which is the largest $Q$ of the range in
consideration, is also shown
in the table. It is chosen so that the ratio $P(Q_{\rm Max}) / P(0) \approx
10^{-20}$ in the weak couplings.
(iii) Between the strong and weak couplings ($2.0\lsim\beta\lsim 2.75$) the
fittings according to the
polynomial turn out to be very poor ( $\chi^2/d.o.f. \approx 250 $ ).
It may indicate the existence of a transitive region between the Gaussian
and non-Gaussian
regions.
(iv) Apart from this region, each of the coefficients change smoothly
from the strong to weak coupling
regions as shown in Table I.\par
Here we discuss the volume dependence.
In the strong coupling regions, $P(Q)$ is fitted very well by Gaussian for all
values
of $V$
$$
P(Q) \propto \exp \left( - \kappa_V(\beta) Q^2 \right).
$$
where the coefficient $\kappa_V(\beta) (=a_2)$ depends on $\beta$ and $V$.
Fig.2 shows $\log \kappa_V(\beta)$ vs. $\log V$ for a fixed $\beta(=0.5)$.
We see that $\kappa_V(\beta)$ is clearly proportional to $1/V$
$$
\kappa_V(\beta) = C(\beta)/V.
$$
This $1/V$-dependence of the Gaussian behavior determines the phase
structure of
the strong coupling region.
This will be discussed in detail numerically in \S 3.2 and
analytically in \S 4.\par
The proportionality constant $C$ depends on $\beta$.
As $\beta$ becomes large, $C(\beta)$ monotonically increases;
$C(\beta=0.0)=10.6$, $C(\beta=0.5)=12.3$, $C(\beta=1.0)=15.5$.
\par
Fig.3. shows the volume-dependence of $P(Q)$ for $L= 24$, 36, 48, and 72
in the weak coupling regions ($\beta=3.0$).
We do not find the $1/V$-law as in the strong coupling regions, but
a clear volume dependence is observed. It causes the different behavior of
$F(\theta)$ from that in the strong coupling regions.
\par
\bigskip
\subsec{Free energy and expectation value of topological charge }
\smallskip
Partition function $Z(\theta)$ as a function of $\theta$ is given
by $\four$ from $P(Q)$.
The free energy is
\eqn\f{
F(\theta) = -{1 \over V} \log Z(\theta).
}
In general, the $n$-th order of the moment is given by the derivatives of
$F(\theta)$
\eqn\q{
\langle Q^n \rangle_\theta = -(-i)^n {d^n F(\theta) \over d\theta^n.}
}
\par
In the strong coupling region, we have seen the Gaussian behavior of $P(Q)$,
and
the $1/V$-law appears to hold up to $L=72$.
It is natural to expect that this behavior persists to $V \rightarrow
\infty$.
Let us look at how the $1/V$-law affects $F(\theta)$ and
$\langle Q \rangle_\theta$.
By putting $C(\beta)=12.3$ in
$P(Q) \propto \exp \left[ - \left( C(\beta)/V \right) Q^2 \right] $
for $\beta=0.5$,
we calculate $F(\theta)$ and $\langle Q \rangle_\theta$ from $\f$ and $\q$.
Fig's 4 and 5 show their volume dependence. As $V$ is increased, $F(\theta)$
very rapidly (already at $L=6$) approaches the quadratic form in $\theta$
from below.
Its first moment $\langle Q \rangle_\theta$ develops a peak near
$\theta=\pi$, and the position of the peak quickly approaches $\pi$ as $V$
increases.
The jump in $\langle Q \rangle_\theta$ would arise at $\theta = \pi$
as $V \rightarrow \infty$.
It indicates the first order phase transition at $\theta=\pi$.\par
In the weak coupling regions, on the other hand, we see the different behavior.
Fig.6 shows $F(\theta)$ at $\beta=3.0$.
For $\theta \lsim \pi/2$, $F(\theta)$ is volume independent, while
for $\theta \gsim \pi/2$, the clear volume dependence appears, where
$F(\theta)$ decreases as $V \rightarrow $ large unlike in the strong coupling
case.
We have checked that the result for $L=20$ agrees with that in ref.$\BDSL$
within errors.
The expectation value $\langle Q \rangle_\theta$ is shown in Fig.7.
The singular behavior at $\theta=\pi$, which was seen
in the small $\beta$ region, disappears. The peak gets round and
its locus moves towards small $\theta$ as $V$ increases,
which is opposite to Fig.5. \par
We should make a remark about the errors in the figures.
As a general tendency,
larger errors arise for
larger volume and/or for $\theta \approx \pi$.
It is associated with the algorithm to calculate $Z(\theta)$, $\four$, in which
$e^{i \theta Q} \approx (-1)^Q $ for $\theta=\pi$ yields large cancellation
for slowly falling $P(Q)$ (the behavior at large $V$) in the summation.
It causes large errors of the observables due to the denominator in $\evq$.
This is just the same as the so called sign problem
\ref\LGSWSS{E.~Y.~Loh~Jr., J.~E. Gubernatis, R.~T. Scalettar, S.~R.~White,
D.~J.~Scalapino
and R.~L.~Sugar, \PRB{41}, 9301 (1990).}
which is notorious
in the quantum Monte Carlo simulations applied to systems of strongly
correlated electrons.
\par
\newsec{Gaussian distribution and the partition function zeros }
In the previous sections, we have seen that $P(Q)$ is Gaussian in
small $\beta$ region.
In this section we shall look into the detail of its consequence by
paying attention to the partition function zeros in the complex $\zeta$
plane ($\zeta=e^{i \theta}$).
Study of the partition function zeros is regarded as an alternative
to investigate the critical phenomena. The zeros accumulate in infinite
volume limit to the critical point, and how fast they approach the point as $V$
increases tells the order of the phase transition
\ref\IPZ
{C.~Itzykson, R.~B.~Pearson and J.~B.~Zuber \NP{220[FS8]}, 415 (1983).}
\ref\FB{M.~E.~Fisher and A.~N.~Berker, \PRB{26}, 2507 (1982).}.
If Gaussian behavior $P(Q) \propto \exp [- \left( C(\beta)/V \right) Q^2]$
persists to
infinite volume limit, the partition function
is expressed by the third elliptic theta function
\eqn\ell{
\vartheta_3(\nu,\tau)=\sum_{Q=-\infty}^\infty p^{Q^2} \zeta^{Q}
}
as
$$
Z(\theta) \propto \vartheta_3(\nu,\tau),
$$
where $p=\exp[-C(\beta)/V] \equiv \exp(i \pi \tau)$
and $\zeta=e^{i\theta} \equiv \exp(i 2 \pi \nu)$.
In order to look for the partition function zeros in the
complex $\zeta$ plane, it is convenient
to use infinite product expansion of $\vartheta_3$
\eqn\ipe{
\vartheta_3(\nu,\tau)=\prod_{m=1}^\infty (1-p^{2 m})
\prod_{n=1}^\infty [(1+p^{2 n-1}\zeta)(1+p^{2 n-1}\zeta^{-1})].
}
Zeros of $Z(\theta)$ are all found easily on the negative real axis of the
complex $\zeta$ plane as
\eqn\zer{
\zeta=-e^{-(2 n-1)C/V}, -e^{+(2 n-1)C/V}
}
for $n$ = 1, 2, $\cdots$, $\infty$.
In the complex $\theta$ plane, equivalently, these zeros are located at
$$
\theta=\pi\pm i (2 n-1) C/V.
$$
It thus follows that the $1/V$-law approaching the critical point
$\theta_c=\pi$ indicates the first order phase transition $\IPZ$ $\FB$.\par
An alternative to the above way of looking is to use the
Poisson sum formula to the sum \four.
\eqn\poi{Z(\theta) \propto \sum_{Q=-\infty}^\infty e^{-C Q^2/V} e^{i\theta Q}
=\sqrt{V\pi/C} \sum_{n=-\infty}^\infty e^{-(\theta -2\pi n)^2 V/4 C}.
}
For $V \gg 1$ and near $\theta=\pi$, the sum on the right is well approximated
by two terms ( $n=0$ and 1 ),
\eqn\pois{ (4.4)
\approx\sqrt{V\pi/C}\left[ e^{-\theta^2 V /4 C}
+ e^{-(\theta-2\pi)^2 V/4 C} \right].
}
It follows that the partition function has infinite zeros at
$\theta=\pi + i (2 n+1)C/V$, where $n$ is integer.
Again the $1/V$-law means the existence of the first order phase transition.
This result is in complete agreement with that from $\vartheta_3$ function
discussed above.
To see to what extent the approximation $\pois$ is good, we compare the
resulting $F(\theta)$ and $\langle Q \rangle_\theta$ from $\pois$ with
those of
Monte Carlo simulations. They agree each other.
\bigskip
\newsec{Conclusions and discussion}
We have seen that $P(Q)$ is Gaussian in the small $\beta$ region.
As shown in the last section, it leads to the first order phase
transition. This behavior is very much like the $d=2$ U(1) gauge model
with $\theta$ term \W, where $P(Q)$ is Gaussian for all values of
the coupling constant \ref \HITY{A.~S.~Hassan, M.~Imachi, N.~Tsuzuki and
H.~Yoneyama,
``Character Expansion, Zeros of partition function and
$\theta$ term in U(1) gauge theory", preprint KYUSHU-HET-25, SAGA-HE-86,
hep-lat/9508011.}. There the analytic form of $P(Q)$ is given.
It may also be interesting to study the $CP^1$ model from the
renormalization group point of view, which might show the singular
behaviors of the renormalization group flows similar to the U(1) case \HIY.
\par
In large $\beta$ region, on the other hand, $P(Q)$ differs from the
Gaussian behavior. Consequently, the free energy $F(\theta)$ and the moment
$\langle Q \rangle_\theta$ show the quite different behaviors from those in the
small $\beta$
regions. The signal of the first order transition disappears. To understand
those behaviors,
It would be helpful to consider the dilute gas approximation,
where instantons of charge $Q=\pm 1$ are randomly distributed.
Let us assume that the probability distribution $P_n$ ( $P_{\overline n}$
), in which $n$ instantons (${\overline n}$ anti-instantons) generate,
obeys the Poisson distribution
$P_n=\lambda^n e^{-\lambda}/n! ( P_{\overline n}=\lambda^{\overline n}
e^{-\lambda}/{\overline n}! ) $. The topological charge distribution function
$P(Q)$ is given by the modified Bessel's function as
$P(Q)=e^{-\lambda} I_Q(\lambda)$, where $\lambda$ is
average number of instantons (anti-instantons).
For $\lambda \gg 1$, $I_Q(\lambda)$ is approximated by $\exp(-Q^2/2\lambda)$.
The $\lambda$ can then be identified as $V/2C$,
which is natural since the average number is proportional to
the volume $V$. As $\beta$ increases, $C(\beta)$ increases
(section 3), that is, the average number of instantons decreases;
as $\beta \rightarrow \infty$ (zero temperature limit), the configurations
vary slowly so that the configurations with large $Q$ are unlikely to
contribute to the partition function. In large $\beta$ region, the behavior of
$I_Q(\lambda)$ as a function of $Q$ is qualitatively the same with the
result of the simulations.
Precisely speaking, however, they are different, and actually the difference
is attributed to
the asymptotic scaling of the topological susceptibility in ref. \BDSL.
\par
It is expected from the Haldane conjecture
that the second order phase transition would occur at $\theta=\pi$.
We, however, seem to fail confirming it. The first order phase transition in
small $\beta$ region would have to mutate to the second order one at some
$\beta$,
if it occurred.
In the large $\beta$ region, as discussed in section 3, the volume dependence
of the
results is large in the interesting region of $\theta$, and the statistical
errors
mask the nature.
For $L=72$, the maximal lattice extension of our study, $F(\theta)$ still
changes considerably and
gets very large errors for $\theta \gsim \pi/2$. Consequently, so do its
moments for
a wider range of $\theta$.
This is due to the large correlation length in the large $\beta$ region, and
the finite size effect is not negligible.
The large fluctuations come from the same origin as the so called
sign problem \LGSWSS, which arises in the strongly correlated electronic system
in the condensed matter physics.
In order to circumvent the problem, we must address the issue of the lattice
effect.
It is worthwhile to pursue the issue treated in the present paper
from the the improved point of view
such as the perfect action
$\ref\HN{P.~Hasenfratz and F.~Niedermayer, \NP{414}, 785 (1994).}$
$\ref\DFP{M.~D'Elia, F.~Farchioni and A.~Papa,
``Scaling and topology in the 2-$d$ O(3)-$\sigma$ model on the lattice
with the perfect action", preprint IFUP-TH 23/95, hep-lat/9505004.}$.
Recently, the second order phase transition has been found numerically by
formulating
the model in terms of clusters with fractional topological charge $\pm 1/2$
\BPW.
\par
Some numerical studies of the $CP^{N-1}$ model with $N > 2$
have been done without $\Pisa$ and with the $\theta$ term \SC.
In the latter case for $CP^3$, interestingly, the first order transition is
observed at finite $\theta$ which is smaller than $\pi$ \SC. \par
\vskip1cm
\centerline {\bf Acknowledgment}
We are grateful to the colleagues for useful discussion.
We also wish to thank S.~Tominaga for discussion on the algorithm.
The numerical simulations were performed on the computer Facom M-1800/20 at
RCNP,
Osaka University.
This work is supported in part by a Grant-in-Aid for Scientific Research
from the Japanese Ministry of Education, Science and Culture (No.07640417).
One of the authors (A.~S.~H.) is grateful for the scholarship from the
Japanese Government.
\vfill\eject
\centerline{{\bf Appendix }}
\smallskip
In this appendix, we discuss briefly
the error analysis when ``set method" and ``trial
distribution method" are used.\par
We consider first the simple case where a single set is adopted and,
as trial distribution, $P_t(Q) = 1$.
It is known that the counts in the histogram method essentially
obeys multinomial distribution and
that the error of counts at $Q$ ($count(Q)$) is estimated by the variance of
the
distribution \BBCS.
For each $Q$, the variance is
$$
\sigma^2(Q) = N \cdot {count(Q) \over N} \left( 1 - {count(Q) \over N} \right),
$$
where $N$ is the total counts. Therefore $P(Q)$ is estimated by
$$
P(Q) = {count(Q) \over N} \pm \delta P(Q),
$$
where $\delta P(Q) = \sigma (Q) / N$.
The relative error ($\delta P / P$) at large $Q$
is given by
$$
{\delta P(Q) \over P(Q)} \approx {\sigma (Q) \over count(Q) }
= {1 \over \sqrt{ count(Q)} }. \eqno (A.1)
$$
It could become very large at large $Q$ when $P(Q)$ is rapidly decreasing
function
of $Q$.\par
When the above two methods are adopted, the relative error decreases as
follows.
The trial distribution method makes $count(Q)$ almost independent of $Q$.
The variance $\sigma(Q)$ also becomes
almost constant at each $Q$. Accordingly, $P(Q)$ is given by
$$
P(Q) = P_t(Q) \left( count(Q) \pm \sigma(Q) \right),
$$
which leads to the relative errors at any $Q$
$$
{\delta P(Q) \over P(Q)} = {\sigma \over count } \approx {\rm constant}.
$$
This is quite an improvement compared to (A.1).
When the set method is further used, the constant errors do not
propagate over different sets \W.
\vfill\eject
\listrefs
\vfill\eject
\centerline{\bf Table caption}
\bigskip
\item{Table I.} The results of chi-square-fitting to $\log P(Q)$ in terms of
the
polynomial $\sum_n a_n Q^n $ for various $\beta$. Fittings are performed to
the data
in the range from $Q=0 $ to $Q_{\rm Max}$. The resulting $\chi^2/d.o.f.$'s
are also listed.
For the data $\beta=0.0$, 0.5 and 1.0, Gaussian fitting is performed.
\par
\vskip 1.5cm
\centerline{\bf Figure captions}
\bigskip
\item{Figure 1.} The topological charge distribution $P(Q)$ vs. $Q^2$ for
$\beta=0.0$ to 3.5. The lattice size is $L=24$.
The data only for $Q \leq 21$ are plotted. The lines are shown for the
guide of eyes.\par
\item{Figure 2.}$\log a_2 ( = \log \kappa_V )$ vs. $\log V$. $\beta = 0.5$.
The $1/V$ behavior is clearly seen. \par
\item{Figure 3.}$P(Q)$ vs. $Q^2$ for $\beta=3.0$.
The lattice size $L$ is taken to be 24, 36, 48 and 72.\par
\item{Figure 4.}Free energy $F(\theta)$ for $\beta=0.5$.
Lines are shown for $V=$ 16, 25 and 36
in order from below.\par
\item{Figure 5.}The expectation value of the topological charge $\langle Q
\rangle_\theta $
for $\beta=0.5$. Lines are shown for $V=$ 16, 25 and 36
in order from below. The peak of the curve becomes sharper quickly as $\theta
\rightarrow \pi$.\par
\item{Figure 6.}$F(\theta)$ for $\beta=3.0$. $L$ is chosen to be 24
(square),
36 (triangle), and 48 (circle).
Values of $F(\theta)$ are plotted based on the parameters $a_n$ obtained by the
fittings explained in the text. The parameters $a_n$ for $L=24$ are shown in
Table I.
Those for $L=36$ and 48 are obtained in the same process as for $L=24$.
The lines are shown for the guide of eyes.
The volume dependence appears clearly at $\theta \gsim \pi/2$. \par
\item{Figure 7.}$\langle Q \rangle_\theta $ for $\beta=3.0$. $L$ is the same as
those
in Fig.6. Error bars for the data of $L=48$ are not drawn because they are too
large.\par
\vfill\eject
\eject
$$\vbox{
\vskip 5cm
\epsfysize=.950\hsize
\epsfbox{tab.ps}
}$$
\smallskip
\vfill
\eject
$$\vbox{
\vskip 2cm
\centerline{Figure 1}
\bigskip
\epsfysize=1.1\hsize
\hskip.05cm
\epsfbox{f1.ps}
}$$
\smallskip
\vfill
\eject
$$\vbox{
\vskip 3cm
\centerline{Figure 2}
\bigskip
\hskip1.3cm
\epsfysize=.6\hsize
\epsfbox{f2.ps}
}$$
\smallskip
\vfill
\eject
$$\vbox{
\vskip 2cm
\centerline{Figure 3}
\bigskip
\epsfysize=1.1\hsize
\hskip.05cm
\epsfbox{f3.ps}
}$$
\smallskip
\vfill
\eject
$$\vbox{
\vskip 1.7cm
\centerline{Figure 4}
\bigskip
\hskip.2cm
\epsfysize=.5\hsize
\epsfbox{f4.ps}
\vskip 1cm
\centerline{Figure 5}
\bigskip
\hskip.2cm
\epsfysize=.5\hsize
\epsfbox{f5.ps}
}$$
\smallskip
\vfill
\eject
$$\vbox{
\vskip 2cm
\centerline{Figure 6}
\bigskip
\epsfysize=.9\hsize
\hskip.05cm
\epsfbox{f6.ps}
}$$
\smallskip
\vfill
\eject
$$\vbox{
\vskip 2cm
\centerline{Figure 7}
\bigskip
\epsfysize=.9\hsize
\hskip.05cm
\epsfbox{f7.ps}
}$$
\smallskip
\vfill
\eject
\bye
|
\section{Introduction}
In a recent series of papers \cite{bs1}-\cite{bmmt} the role of Pauli
exclusion principle to explain the experimental data on unpolarized and
polarized structure functions of the nucleons has been studied.
The relevance of the quantum statistics on the parton distribution is
supported by several phenomenological observations. The most relevant
phenomenon is certainly the measurement of a defect in the Gottfried sum
rule \cite{gott, NMC}. It can be explained in terms of a Pauli blocking
effect on the production of $u$ sea quark with respect to $d$ sea quark,
which yields a flavour asymmetry between $\bar{u}$ and $\bar{d}$ in the
proton. Another interesting observation is a relationship which seems to
occur between the shapes of the quark parton distributions and their first
momenta, which is the typical characteristic of Fermi--Dirac distribution
functions.
These considerations naturally suggest to use quantum statistically inspired
parameterizations for the parton distribution functions. In this description,
the independent variable which plays the role of the energy is the Bjorken
variable $x$, and the distributions assumed will be of the Fermi--Dirac kind
for quarks, and Bose--Einstein for gluons. Interestingly, to properly describe
the low $x$ behaviour of the structure functions, a {\it liquid} unpolarized
component dominating the very low $x$ region has to be added. It does not
affect the quark parton model sum rules (QPMSR) but it is necessary to
reproduce the antiquark distribution at low $x$.
The paper is organized as follow, in Section 2 we summarize
the phenomenological motivations behind our description for the parton
distributions. In Section 3 the distribution function parameterizations
are shown in detail and discussed in connection with a possible
physical interpretation. Section 4 deals with the results of a fitting
procedure performed to get the free parameters of the distribution functions.
The theoretical predictions are shown in comparison with the experimental
data and the results for QPMSR are also discussed. In Section 5
we give our conclusions and remarks.
\section{Evidence for quantum statistical effects in parton distributions}
As already stated, an experimental evidence for a central role played by the
Pauli principle in the physics of nucleon is the defect in the Gottfried sum
rule \cite{gott}
\begin{equation}
I_{G} = \int_{0}^{1} { 1\over x}\left[{F_{2}^{p}}(x) - {F_{2}^{n}}(x) \right]~dx = {
1\over 3} ( u + \bar{u} -d -\bar{d} ) ~~~,
\label{9}
\end{equation}
that for $SU(2)_{I}$ invariant sea quark distributions ($\bar{d} = \bar{u}$)
gives $I_G=1/3$. Indeed NMC experiment \cite{NMC} measures for the
l.h.s. of Eq. (\ref{9})
\begin{equation}
I_{G} = 0.235 \pm 0.026~~~,
\label{10}
\end{equation}
implying
\begin{equation}
\bar{d} - \bar{u} = 0.15 \pm 0.04 ~~~~~~~~~~~~ u - d = 0.85 \pm 0.04~~~.
\label{11}
\end{equation}
This experimental result can be explained following the idea of Field and
Feynman \cite{Field}, who suggested that, in the proton, the Pauli principle
depresses the production of $u \bar{u}$ pairs in the proton with respect to
$d \bar{d}$, since it contains two valence $u$ quarks and only one $d$. We
will return on this fact in the following to focus on a possible connection
between the violation of the above sum rule and the Bjorken one \cite{Bj}.
{}From the previous considerations one expects a relevant role played by the
statistics in the whole phenomenology of deep inelastic scattering, and thus
it suggests to look for others typical characteristics of this behaviour.
A peculiar characteristic of a Fermi--Dirac statistical function is certainly
the strong connection between shape and abundance. This is an immediate
consequence of the Pauli exclusion principle which forbids two or more
fermions to have the same quantum numbers and implies that the more abundant
is distribution the broader is in $x$ the associated function.
With the aim to check if this situation occurs in the nucleon structure
let us consider the abundances of valence quarks in the nucleons. As
it is well-known, at $Q^2 =0$ they are connected to the axial couplings
of the baryon octet $F$ and $D$, through the relations
\begin{equation}
u^{\uparrow}_{val} = 1+F~~~~~~~~~~~~u^{\downarrow}_{val} = 1-F~~~,
\label{1}
\end{equation}
\begin{equation}
d^{\uparrow}_{val} = {1+F-D \over 2}~~~~~~~~~~d^{\downarrow}_{val} = {1-F+D \over 2}~~~.
\label{2}
\end{equation}
By using the experimental values obtained by the two bodies strong decays
of hyperons $F=0.464 \pm 0.009$ $D=0.793 \pm 0.009$ \cite{pdg94}, \cite{3},
we get
\begin{equation}
u^{\uparrow}_{val} \simeq { 3 \over 2} \simeq u^{\downarrow}_{val} + d^{\uparrow}_{val} + d^{\downarrow}_{val}~~~,
\label{4}
\end{equation}
which tells us that $u^{\uparrow}_{val}$ is the most abundant parton in the proton,
at least at $Q^2=0$. Moreover, by observing that $F \simeq 1/2$ and
$D \simeq 3/4$ one also gets
\begin{equation}
u^{\downarrow}_{val} \simeq { 1 \over 2} = {d^{\uparrow}_{val}+d^{\downarrow}_{val} \over 2 }~~~.
\label{4b}
\end{equation}
In addition to this, the behaviour at high $x$ of ${F_{2}^{n}}(x)/{F_{2}^{p}}(x)$ \cite{1}, known
since a long time, and the more recent polarization experiments \cite{2},
\cite{2bis}, which show that at high $x$ the partons have spin parallel to
the one of the proton, imply that $u^{\uparrow}(x)$ is the dominating parton
distribution in the proton at high $x$. Thus, to the most abundant
$u^{\uparrow}$ corresponds effectively a broader distribution in the Bjorken
variable $x$.
Eq. (\ref{4b}) has also another interesting implication. In a previous work
we assumed that the parton distributions at a given large $Q^2$ depend on
their first momenta (abundances) computed at $Q^2 =0$ \cite{bs2}
\begin{equation}
p(x) = {\cal F}(x,p_{val})~~~,
\label{5}
\end{equation}
with ${\cal F}$ an increasing function of $p_{val}$ and with a broader shape
for higher values of $p_{val}$. From this assumption and by virtue of
(\ref{4b}) we get
\begin{equation}
u^{\downarrow}(x) = { 1 \over 2} \left[ d^{\uparrow}(x) + d^{\downarrow}(x) \right] = { 1 \over 2}
d(x)~~~,
\label{6}
\end{equation}
which implies
\begin{equation}
\Deltau(x) = u^{\uparrow}(x) - u^{\downarrow}(x) = u(x) - d(x)~~~.
\label{7}
\end{equation}
Note that, Eq. (\ref{7}) connects the contribution of $\Deltau(x)$ to
${g_{1}^{p}}(x)$, with the terms due to up and down quarks in the unpolarized
structure functions of nucleons $F_{2}(x)$ \cite{bs2}
\begin{equation}
x {g_{1}^{p}}(x) \Bigr|_{\Delta u} = {2 \over 3} \left[ {F_{2}^{p}}(x) -{F_{2}^{n}}(x)
\right]_{u+d}~~~.
\label{8}
\end{equation}
This relation should hold in good approximation for the total
quantities $x {g_{1}^{p}}(x)$ and ${F_{2}^{p}}(x) -{F_{2}^{n}}(x)$, since the contribution in
${g_{1}^{p}}(x)$ due to $\Deltad(x)$ is depressed for the twofold reason that
$e_{d}^2 = (1/4) e_{u}^2$ and $\Delta d_{val} \simeq - (1/4)
\Delta u_{val}$. By integrating Eq. (\ref{8}) one thus get a connection
between the spin sum rule and the Gottfried sum rule and in turn a
relation between their possible defects.
\section{Parton distributions as Fermi--Dirac and Bose--Einstein statistical
functions}
The previous considerations on the role played by the Pauli principle
in the nucleon structure suggest to assume Fermi--Dirac distributions
in the variable $x$, at least for large $x$, for the quark partons
\cite{bbmmst}
\begin{equation}
p_\lambda(x) = f(x) \left[\exp\left({ x - \tilde{x}(\lambda) \over
\bar{x}} \right) + 1 \right]^{-1}~~~,
\label{px}
\end{equation}
where the index $\lambda$ denotes the different species of quarks,
characterized by flavour and polarization. In Eq. (\ref{px}), $f(x)$ is a
weight function which accounts for the energy level density, and because it
is connected to the nonperturbative aspect of QCD results independent of
flavours and polarization. The {\it universal} parameter $\bar{x}$ represents
the {\it temperature} for the system, whereas $\tilde{x}(\lambda)$ stands the
{\it thermodynamical potential} of the parton $\lambda$.
The expression chosen for $f(x)$ is inspired by the expected power behaviour
at $x=0$, and by the obvious kinematical cut
which forces the function to vanish at $x=1$. In order to satisfy
these constraints we assume for simplicity a power low dependence on $x$
\begin{equation}
f(x)= A~ x^{\alpha} (1 -x)^{\beta}~~~.
\label{fx}
\end{equation}
Indeed, the assumption that the form given in Eq. (\ref{px}) for the quark
distribution functions, which requires different thermodynamical potentials
in order to describe the experimental data, is valid in the low $x$
limit as well has at least two unpleasant features. Firstly, one gets in the
nonperturbative region different behaviour for the different parton
distributions, where on the contrary one would expect an {\it universal}
dependence on $x$. Moreover, the power dependence on $x$ of Eq. (\ref{fx}),
fitted by the experimental data mostly placed at the large $x$, is not
suitable to reproduce the more divergent contribution expected at low $x$.
This most divergent part does not contribute to QPMSR as the ones given by
Gottfried and Bjorken \cite{Bj} with $I=1$ quantum numbers exchanged.
To this aim we add to (\ref{px}) an unpolarized component, which we call
{\it liquid} to stress the possibility that it is connected to the presence,
at low $x$, of a new phase in the quark-gluon plasma due to the highly
nonperturbative QCD regime
\begin{equation}
p_\lambda(x) = {A_{L} \over 2}~x^{\alpha_{L}} (1-x)^{\beta_L} +
A~ x^{\alpha} (1 -x)^{\beta} \left[\exp\left({ x - \tilde{x}(\lambda)
\over \bar{x}} \right) + 1 \right]^{-1}~~~.
\label{ptx}
\end{equation}
In the fitting procedure we take as free parameters, apart from the constants
involved in $f(x)$ and in the {\it liquid} component of (\ref{ptx}), the
temperature $\bar{x}$ and the $\tilde{x}$ for $u^{\uparrow(\downarrow)}$,
$d^{\uparrow(\downarrow)}$, $\bar{u}$ and $\bar{d}$ (the latters are assumed
not polarized). We tried initially to introduce spin-dependent $\tilde{x}$'s
also for the $\bar{q}$'s and to test the relationship
\begin{equation}
\Delta \bar{u}(x) = \bar{u}(x) - \bar{d}(x)~~~,
\label{1a}
\end{equation}
assumed in previous works \cite{bs2, bos1}, but
unfortunately, we found practically the same $\chi^2$ with negative
and positive values for $\Delta\bar{u}(x)$ and/or $\Delta\bar{d}(x)$.
Hence, for not loosing predictivity in the fit procedure we have
assumed unpolarized antiquarks.
As far as the strange quarks are concerned, we assume for simplicity
unpolarized distribution functions given by the empirical expression
\begin{equation}
s = \bar{s} = {\bar{u}+\bar{d} \over 4}~~~,
\label{sx}
\end{equation}
which experimentally is very well satisfied.
Analogously, for the gluons, if we neglect their polarization , the
bosonic statistic suggests the consider, at large $x$, the distribution
function
\begin{equation}
G(x) = {16 \over 3} A~ x^{\alpha} (1 -x)^{\beta}
\left[\exp\left({ x - \tilde{x}_{G} \over \bar{x}} \right) - 1
\right]^{-1}~~~,
\label{gx}
\end{equation}
where the factor $16/3$ is just the product of $2$ ($S_{z}(G) = \pm 1$)
times $8/3$, the ratio of the colour degeneracies for gluons and
quarks.
\section{Discussion of the results}
By assuming for the parton distributions Eqs. (\ref{ptx}) and (\ref{gx}),
we fit the distribution parameters from the experimental data for
${F_{2}^{p}}(x) -{F_{2}^{n}}(x)$ \cite{NMC}, $xF_{3}(x)$ \cite{12}, $x{g_{1}^{p}}(x)$ \cite{2, 2bis} and
$x{g_{1}^{n}}(x)$ \cite{16}, which do not receive contributions from the {\it liquid}
component, and from ${F_{2}^{n}}(x)/{F_{2}^{p}}(x)$ \cite{NMC} and $x\bar{q}(x)$ \cite{9}.
The avaliable experimental data on deep inelastic scattering observables
correspond in general to different values of $Q^2$. This would suggest,
in order to use the data to determine the distribution parameters, which in
general will depend on $Q^2$, to apply the evolution equations to lead all
the experimental results to the same $Q^2$. In our analysis we have neglected
this $Q^2$ dependence of the distribution parameters, since we expect from
the evolution equations a smooth logarithmic dependence.
As far as the polarized
distributions is concerned, in fact, the expected $Q^2$ dependance \cite{bos1}
results to be smaller than the experimental errors, and thus our analysis
is slightly affected by neglecting this dependance.
Indeed, the data on unpolarized nucleons structure functions are at
$Q^2 = 4~GeV^2$ \cite{NMC}, the neutrino data at $Q^2 = 3~GeV^2$ \cite{12},
and $\bar{q}$ measures are performed at $Q^2 = 3~GeV^2$ and $5~GeV^2$
\cite{9} and differ at small $x$, while our curve is intermediate
between the two sets of
data. The data on $g_1^n(x)$ are at $Q^2 = 2~GeV^2$ \cite{16},
whereas $g_1^p(x)$ is measured at $Q^2 = 10~GeV^2$ by SMC \cite{2} and
at $Q^2 = 3~GeV^2$ by E143 \cite{2bis}; despite some narrowing of
the distribution at higher $Q^2$ showing up in the data, the values of
$I_p$ are in good agreement.
In Table 1 we report the parameters found in the present analysis
\cite{bmmt}and compare them with the results of a previous fit
(without liquid) \cite{bbmmst}, and with the ones by Bourrely and
Soffer \cite{bos1} found on similar principles, but with several different
assumptions. In the Figures 1.-6., the predictions for the nucleon structure
${F_{2}^{p}}(x) -{F_{2}^{n}}(x)$, ${F_{2}^{n}}(x)/{F_{2}^{p}}(x)$, $x{g_{1}^{p}}(x)$, $x{g_{1}^{n}}(x)$, $xF_{3}(x)$ and $x\bar{q}(x)$
are shown, respectively, and compared with the experimental data.
The parton distributions found in \cite{bmmt} are described in Figure 7.
Since the total
momentum carried by fermion partons is $53\%$, we get $\tilde{x}_G=-1/15$
by requiring that the gluons carry out the remaining part of the
proton momentum. In Ref. \cite{bmmt},
the gluon distribution is compared with the information
found on them in CDHSW \cite{CDHSW}, SLAC$+$BCDMS \cite{SLAC&BCDMS} and
in NMC \cite{NMCgluon} experiments at $Q^2 = 20~GeV^2$. The
agreement is fair for $x > .1$, while the fast increase at small $x$,
confirmed also from the data at very small $x$ at Hera \cite{H1gluon},
confirms that a liquid component is needed also for gluons. The excess
at high $x$ of our curve with respect to experiment may be, at least
in part, explained by the expected narrowing of the distribution from
$Q^2 = 4~GeV^2$, where we fit the unpolarized distributions,
to $Q^2 = 20~GeV^2$.
The inclusion of the {\it liquid} term and the extension of our fit to
the precise experimental results on neutrinos has brought to substantial
changes in the parameters \cite{bmmt} with respect to the previous work
\cite{bbmmst}.
The low $x$ behaviour of $f(x)$ become smoother ($\simeq x^{-.203 \pm .013
}$ instead of $x^{-0.85}$), but this is easily understood since the
previous behaviour was a compromise between the smooth {\it gas}
component and the rapidly changing {\it liquid} one to reproduce the
behaviour of $\bar{q}(x)$. The {\it liquid} component, relevant only
at small $x$, carries only $.6\%$ of parton momentum and its behaviour
$\sim x^{-1.19}$, similar to the result found in \cite{capella},
is less singular than the one, suggested in the
framework of the multipherial approach to deep-inelastic scattering,
proportional to $\sim x^{-1.5}$ \cite{15}. The parameter
$\tilde{x}(u^{\uparrow})$ took the highest value allowed by us (1.),
since the factor in $f(x)$,
$(1-x)^{2.34}$, is taking care to decrease $u^{\uparrow}(x)$ at high $x$. The
temperature $\bar{x}$ is larger than
the previous one and the one found by Bourrely and Soffer \cite{bos1}. Instead
$\tilde{x}(u^{\downarrow})$ is slightly smaller than the previous
determination \cite{bbmmst} and about half the value found in \cite{bos1},
where $f(x)$ is different for $u^{\uparrow}$ and $u^{\downarrow}$.
The ratio $r = u^{\downarrow}(x)/d(x)$ varies in the
narrow range $(.546,.564)$ in fair agreement with the constant value
$1 -F = .536 \pm .009$ assumed in \cite{bbmmst} and
slightly larger than the value $1/2$ taken in \cite{bs2} and \cite{bos1}.
The central value found for the first moment of $\bar{u}_{gas}(x)$, $.03$,
is smaller than $\bar{d}_{gas}(x)/2$, $.08$, while Eq. (\ref{1a})
implies $\bar{u}(x) \geq \bar{d}(x)/2$. However, the large upper error on
$\bar{u}_{gas}$ and the uncertainty
in disentangling the gas and liquid contributions for the $\bar{q}$'s do not
allow to reach a definite conclusion about the validity of Eq. (\ref{1a}).
Finally, we can compare the predictions of \cite{bmmt}
with the measured asymmetry for Drell-Yan
production of muons at $y=0$ in $pp$ and $pn$ reactions
\begin{equation}
A_{DY} = { d \sigma_{pp}/dy - d \sigma_{pn}/dy \over
d \sigma_{pp}/dy + d \sigma_{pn}/dy}~~~,
\label{ady}
\end{equation}
which at rapidity $y=0$ reads
\begin{equation}
A_{DY} = { (\lambda_s(x) -1) (4 \lambda(x) -1) +
(\lambda(x) -1) (4 \lambda_s(x) -1) \over
(\lambda_s(x) +1) (4 \lambda(x) +1) +
(\lambda(x) +1) (4 \lambda_s(x) +1) }~~~,
\label{ady1}
\end{equation}
where $\lambda_s(x)=\bar{u}(x)/\bar{d}(x)$ and
$\lambda(x)=u(x)/d(x)$. At $x=.18$ we have $\lambda_s(.18)=.454$
and $\lambda(.18)=1.748$ giving rise to $A_{DY}(.18)=-.138$ in
fair agreement with the experimental result $-.09 \pm .02 \pm .025$
\cite{NA51}.\\
The behaviour of $A_{DY}(x)$ is plotted in Figure 8 together with
the experimental point measured by NA51 collaboration.
\section{Conclusions}
We compared with data the quark-parton distributions given by the sum of
Fermi--Dirac functions and of a term not contributing to the QPMSR relevant at
small $x$. We obtain a fair description for the unpolarized and polarized
structure functions of the nucleons as well as for the $F_3(x)$ precisely
measured in (anti)neutrino induced deep-inelastic reactions and for $\bar{q}$
total distribution. The conjectures of previous works on $d$ distributions are
well confirmed by the values chosen for their thermodynamical potentials. As
long as the implications for QPMSR the values found for the first momenta of
the various parton species give l.h.s.'s consistent with experiment. For the
fundamental issue of the Bjorken sum rule, as advocated in previous works
\cite{bs1, bs4} and \cite{bbmmst}, we get
\begin{eqnarray}
\Delta u & \simeq & u - d + 2F-1~~~,\label{37}\\
\Delta d & \geq & F- D~~~,\label{38}
\end{eqnarray}
to confirm the suspicion of a violation of Bjorken sum rule related to
the defect in the Gottfried sum rule.
A word of caution is welcome for our conclusions on the violation of
Bj sum rule, since we did not include the effect of QCD corrections
in relating the quark parton distributions to the structure functions.
Also we assumed no polarization for $\bar{q}$, being unable to get a
reliable evaluation of $\Delta \bar{q}$ with the present precision
for the polarized structure functions at
small $x$. Indeed our description of $g_1^p(x)$ and $g_1^n(x)$
is good in terms of $\Delta u(x)$ and $\Delta d(x)$, but our
prediction is smaller than the central values of the three lowest $x$
values measured by SMC.
\newpage
|
\section{Introduction}
The radiatively-driven winds of early-type (O, B, Wolf-Rayet) stars
are observed to vary on time scales ranging from hours to years.
In addition, as in the highly complex solar wind, the mass outflows
from hot stars are presumably not spherically symmetric.
There are many physical mechanisms that can lead to wind structure
and variability, and it is useful to distinguish between
(1) {\em small-scale} stochastic fluctuations, intrinsic to the
wind itself, and
(2) {\em large-scale} quasi-regular variability, induced by changes
in the underlying star.
In the former category is the the shocked structure arising from
the strong instability of the line-driving mechanism
(Rybicki \markcite{R87}1987; Owocki \markcite{O92}1992),
which may explain black troughs in saturated UV P~Cygni lines in
OB~stars (Lucy \markcite{L82}1982; Puls, Owocki, \& Fullerton
\markcite{POF}1993, hereafter POF),
shock-heated X-ray emission (Cooper \& Owocki \markcite{CO94}1994),
and moving ``bumps'' in Wolf-Rayet optical emission lines
(Robert \markcite{R94}1994).
On the other hand, the larger-scale structure could be attributed to
the dynamical effects of rotation,
magnetic fields, or nonradial pulsations, which
may produce the recurring discrete absorption components (DACs)
and blue-edge variability observed in ultraviolet P~Cygni lines.
There has been much progress in using radiation hydrodynamics to model
the small-scale intrinsic wind instability, but
considerably less attention has been given to the problem of how
line-driven winds respond to larger star-induced
variations.
This paper reports on initial results from radiation hydrodynamics
models of winds affected by such laterally coherent and
rotationally modulated perturbations.
Large-scale wind structure in hot stars is inferred most directly
from time variability in
the blueshifted absorption troughs of UV P~Cygni profiles.
The most conspicuous variations are the DACs, which
appear as narrow and localized optical depth enhancements
in unsaturated lines,
in some stars even dominating the ``mean wind'' absorption.
DACs are present in a majority of
O-star (Howarth \& Prinja \markcite{HP89}1989) and
Be-star (Grady, Bjorkman, \& Snow \markcite{Ge87}1987) winds,
and are typically seen to accelerate to the blue wing of the profile
over a few days, becoming narrower as they approach an asymptotic
velocity.
Prinja \markcite{P88}(1988) and Henrichs, Kaper, \& Zwarthoed
\markcite{HKZ}(1988) found an
apparent correlation between both the recurrence and acceleration
time scales of DACs
(typically of the same order as each other)
and the projected rotational velocity of the star,
$V_{\mbox{\footnotesize eq}} \sin i$.
Corresponding and often temporally-correlated variability is seen
in the blue edge fluctuations of
saturated UV P~Cygni lines,
in the low-velocity variability of subordinate-level P~Cygni lines,
and in optical lines such as $\mbox{H}\alpha$ and
\ion{He}{2} $\lambda 4686$,
suggesting a single dynamical phenomenon reaching down to very near
the photosphere
(Henrichs, Kaper, \& Nichols \markcite{HKN}1994).
Attempts to model DACs have been progressively constrained
by better observations.
By studying lines of different ionization species, Lamers, Gathier,
\& Snow \markcite{Le82}(1982)
ruled out the early supposition that DACs might be caused by ionization
gradients in an otherwise spherically symmetric and time steady wind.
The episodic ejection of spherical ``shells'' of increased mass loss
was an often-invoked model for a time, but the lack of both
{\em emission} variability in UV P~Cygni lines
(Prinja \& Howarth \markcite{PH88}1988) and significant infrared
variability (Howarth \markcite{H92}1992)
seems to rule out a spherically-symmetric disturbance.
On the other hand, to produce the observed strong absorption dips,
the structure must be large enough to cover a substantial fraction
of the stellar disk.
This seems to rule out the small-scale wind instability as the source
of most DAC clumpiness,
since global averaging would weaken the observable signature
(Owocki \markcite{O94}1994).
Also, Rybicki, Owocki, \& Castor \markcite{Re90}(1990) showed that
small-scale, lateral velocity perturbations should be
strongly damped, and so should not disrupt the horizontal scale
size set by base variations.
Altogether, these constraints suggest that DACs originate from
moderate size wind structures,
e.g., spatially-localized clouds, streams, or ``blobs.''
Of particular interest is the apparent acceleration rate of DACs.
When compared to the acceleration of the mean wind inferred from
line-driven flow theory and detailed profile fitting,
some (typically weaker) DACs seem to be passively carried along the same
velocity law (see, e.g., Kaper \markcite{K93}1993).
But most strong DACs accelerate much more {\em slowly} (Prinja
\markcite{P94}1994),
suggesting they may not represent a single mass-conserving feature,
but rather might arise from a slowly evolving {\em pattern} or perturbation
through which wind material flows.
The enhanced optical depth could result from either a higher density
or a lower wind velocity gradient (a ``plateau''), or by a combination
of the two (Fullerton \& Owocki \markcite{FO92}1992;
Owocki, Fullerton, \& Puls \markcite{Oe94}1994),
as is found in the dynamical models below (\S~3).
Mullan (\markcite{M84a}1984a, \markcite{M84b}b; \markcite{M86}1986)
proposed that DACs and related phenomena
could arise from ``corotating interaction regions'' (CIRs) analogous
to those commonly observed {\em in situ} in the solar wind.
In the solar corona, regions of open magnetic field cause the
flow from coronal holes to be accelerated faster than the mean
ecliptic-plane wind, resulting in colliding fast and slow streams strung
into spiral CIR patterns by rotation (Hundhausen
\markcite{H72}1972; Zirker \markcite{Z77}1977).
These nonlinear interacting streams eventually steepen into
oblique corotating shocks, through which the wind flows nearly radially.
Because hot stars do not have the strong surface convection
and coronae known to exist in the sun, the ``seed'' mechanism for
large-scale azimuthal perturbations may be quite different.
In this paper we do not adhere to any particular model
for these photospheric variations,
but several plausible scenarios have been proposed.
Underhill \& Fahey \markcite{UF84} (1984) and Henrichs et
al.\ \markcite{HKN}(1994) suggested
that small patches of enhanced
magnetic field could exist undetected on early-type stars and
produce corotating wind structure.
Also, nonradial pulsations have been observed in many OB stars, and
have been shown to be able to induce
localized increased mass loss and outward
angular momentum transfer (Castor \markcite{C86}1986; Willson
\markcite{W86}1986; Ando \markcite{A91}1991).
Circumstellar disks exhibit many natural large-scale instabilities,
e.g.\ Okazaki's \markcite{O91}(1991) global one-armed
normal modes, which may be correlated with DAC variability in
Be stars (Telting \& Kaper \markcite{TK94}1994).
Dowling \& Spiegel \markcite{DS90}(1990) discuss the possible existence of
Jupiter-like zonal bands and vortices in the atmospheres of hot
stars, and give order-of-magnitude estimates of the flux
enhancement over a ``Great Red Spot'' type of shear pattern.
Several qualitative attempts have been made to apply the CIR
picture to observations of time variability in early-type stellar
winds, but all have been {\em kinematic} in nature.
Prinja \& Howarth \markcite{PH88}(1988)
fit slowly-accelerating spiral streamlines
to DACs in time series spectra from the O7.5 giant 68~Cyg, and showed
that the narrowing of the absorption feature as it accelerates can
be explained roughly by the decrease in the line-of-sight velocity
gradient of the CIR.
Harmanec \markcite{H91}(1991) extended this analysis and discussed possible
observational signatures of CIRs in other classes of early-type stars.
Rotationally-modulated gas streams or ``spokes'' have been proposed
in models of Be star circumstellar material
(see, e.g., \v{S}tefl et al.\ \markcite{Se95}1995)
and in the time variability of Herbig
Ae star spectra (Catala et al.\ \markcite{Ce91}1991).
Of course, the physics of circumstellar streams in Be and Herbig Ae stars
will most likely be very different from that of corresponding
structures around O stars, and we will focus mainly on the latter.
The principal goal of this paper is to model the
{\em dynamical} effect of radiative
driving on the formation of CIRs in a hot star wind,
and to compute synthetic observational diagnostics to see if, e.g.,
slow DAC-like signatures can be theoretically produced.
Our computational approach is to apply a reasonable parameterization
for a localized ``star spot'' perturbation in the radiation force near
the stellar surface, and allow the wind to respond consistently.
Although the actual photospheric structure perturbing the base of
the wind is likely to be different,
we suspect the characteristic response
of a radiatively-driven medium will be insensitive to the details of
this physical mechanism,
and mainly depend on base changes in the fluid velocity and density.
The remainder of this paper is organized as follows.
We first (\S~2) describe our radiation hydrodynamics code and the
details of the induced azimuthally-dependent force enhancement.
Next (\S~3) we present results for a series of O~star models with
varying rotation rates and wind parameters, and discuss the emergent
corotating structure.
We then (\S~4) compute synthetic UV P~Cygni line profile time series
for the various models in our parameter study.
Finally, a discussion and conclusion section
(\S~5) summarizes our results and
outlines directions for future work.
\section{Numerical Radiation Hydrodynamics}
\subsection{Equations of Hydrodynamics}
The problem of structure formation in a stellar wind is
in general arbitrarily three-dimensional.
Rotation imposes a latitudinal dependence on both the photosphere
and the wind, and large-scale variations at the stellar surface
can impose an arbitrary latitudinal or azimuthal
dependence on the wind.
However, to study how a line-driven wind reponds to rotationally
induced variations within a more tractable, {\em two-dimensional} model,
we confine the simulations here to the {\em equatorial plane,}
where rotation has the strongest impact, and where the flow can be
constrained to a surface of constant colatitude, $\theta = \pi /2$.
(See Pizzo \markcite{P82}1982 for discussion of similar approximations
in modeling the solar wind.)
This assumption naturally suppresses the centrifugal wind compression
effect of Bjorkman \& Cassinelli \markcite{BC93}(1993), which is
only of minor importance for O stars.
The induced ``spot'' variations thus
only have an azimuthal extent, and assumptions about
latitudinal structure are only required when computing
observational diagnostics (\S~4), not the inherent dynamics.
We use a time-dependent numerical hydrodynamics code to evolve a
model of a radiatively-driven wind from a rotating star toward an
equilibrium corotating steady state.
The code, VH-1, was developed by J.\ M.\ Blondin and colleagues
at the University of Virginia,
and uses the piecewise parabolic method (PPM) algorithm developed by
Collela \& Woodward \markcite{CW84}(1984).
VH-1 solves the Lagrangian forms of the equations of hydrodynamics
in the fluid rest frame, and remaps conserved quantities onto an
Eulerian grid at each time step.
The equations to be numerically integrated, written in Eulerian form
using spherical polar coordinates, include the conservation of mass,
\begin{equation}
\frac{\partial \rho}{\partial t} +
\frac{1}{r^2} \frac{\partial}{\partial r}
\left( \rho v_{r} r^{2} \right) +
\frac{1}{r \sin\theta} \frac{\partial}{\partial r}
\left( \rho v_{\phi} \right) \, = \, 0 \,\,\,\, ,
\end{equation}
and the conservation of the $r$ (radial) and $\phi$ (azimuthal)
components of momentum,
\begin{equation}
\frac{\partial v_{r}}{\partial t} +
v_{r} \frac{\partial v_{r}}{\partial r} +
\frac{v_{\phi}}{r \sin\theta} \frac{\partial v_{r}}{\partial \phi}
\, = \, \frac{v_{\phi}^{2}}{r} -
\frac{1}{\rho} \frac{\partial P}{\partial r} +
g_{r}^{\mbox{\footnotesize ext}} \,\,\,\, ,
\end{equation}
\begin{equation}
\frac{\partial v_{\phi}}{\partial t} +
v_{r} \frac{\partial v_{\phi}}{\partial r} +
\frac{v_{\phi}}{r \sin\theta} \frac{\partial v_{\phi}}{\partial \phi}
\, = \, - \frac{v_{r} v_{\phi}}{r} -
\frac{1}{\rho r \sin\theta} \frac{\partial P}{\partial \phi} \,\,\,\, ,
\end{equation}
where $\rho$ is the mass density, $v_{r}$ and $v_{\phi}$ are the $r$
and $\phi$ components of the velocity, and $t$ is the time.
The $\theta$ (latitudinal) component of the momentum conservation
equation is assumed satisfied in the equatorial plane by the trivial
solution $v_{\theta} (\theta = \pi /2) \, = \, 0$, i.e.\ no latitudinal
flow into or out of the computational domain,
with all partial derivatives in the $\theta$ direction
considered negligible.
The code also includes an equation for the conservation of energy,
but in all models presented here this is dominated by rapid radiative
processes, which keep the gas very nearly isothermal
with a constant wind temperature $T$ equal to the stellar effective
temperature $T_{\mbox{\footnotesize eff}}$. We use a perfect gas law
equation of state to evaluate the pressure $P$.
As in Owocki, Cranmer, \& Blondin \markcite{OCB}(1994), the external radial
acceleration here includes gravity and radiative driving by line
scattering, and
\begin{equation}
g_{r}^{\mbox{\footnotesize ext}} \, = \,
- \frac{GM_{\ast} (1 - \Gamma )}{r^2} +
g_{r}^{\mbox{\footnotesize lines}} \,\,\,\, .
\label{eq:geff}
\end{equation}
Here $G$ and $M_{\ast}$ are the gravitational constant and stellar
mass, and $\Gamma$ ($= \kappa_{e} L_{\ast} / 4 \pi GM_{\ast} c$)
is the Eddington factor that accounts for the reduction in effective
gravity by outward radiation pressure on electrons.
We evaluate the
line-scattering acceleration $g_{r}^{\mbox{\footnotesize lines}}$
in the local Sobolev \markcite{S60}(1960) approximation.
This suppresses the wind's strong line-driven instability
(Owocki \& Rybicki \markcite{OR84}1984), as it is not
currently feasible to incorporate this inherently nonlocal effect in
multidimensional hydrodynamic models.
For one-dimensional winds, the Sobolev line force per unit mass can
be parameterized as
\begin{equation}
g_{r}^{\mbox{\footnotesize lines}} \, = \, k f
\left( \frac{1}{\kappa_{e} v_{\mbox{\footnotesize th}} \rho}
\left| \frac{\partial v_{r}}{\partial r} \right| \right)^{\alpha}
\frac{GM_{\ast}\Gamma}{r^2} \,\,\,\, ,
\label{eq:glines}
\end{equation}
where the Castor, Abbott, \& Klein \markcite{CAK}(1975,
hereafter CAK) parameters
$\alpha$ and $k$ are related to the slope and normalization of the
assumed power-law ensemble of lines.
The constant $k$ is defined in terms of the electron scattering
coefficient $\kappa_e$ and a fiducial ion thermal speed
$v_{\mbox{\footnotesize th}}$ (Abbott \markcite{A82}1982).
The stellar radiation field is modeled here by a spherical star
with no limb darkening, and
$f$ is the finite-disk correction factor used
by Friend \& Abbott \markcite{FA86}(1986) and Pauldrach, Puls,
\& Kudritzki \markcite{PPK}(1986).
We ignore the effects of rotational oblateness and
gravity darkening on the full {\em vector} radiative force (Cranmer
\& Owocki \markcite{CO95}1995), which we take to be purely radial.
For simplicity, we also
do not use Abbott's \markcite{A82}(1982) added ionization-balance
parameterization of the line force,
and set his exponent $\delta =0$.
The Sobolev approximation assumes a monotonically accelerating
velocity field, but this is not guaranteed in the time-dependent
simulations presented below.
Winds with nonmonotonic velocities have {\em nonlocal} line forces,
since multiple resonance surfaces can create additional attenuation
of the stellar flux (Rybicki \& Hummer \markcite{RH78}1978;
POF\markcite{POF}).
We do not treat this nonlocal coupling directly, but we have compared
CIR models using an upper limit (unattenuated) and a lower limit
(strongly attenuated) for the line force in multiply-resonant regions
of the wind; we find the dynamics to be quite similar in both limits.
The upper limit, which we use in all models presented below, involves
taking the absolute value of the radial velocity gradient
$| \partial v_{r} / \partial r |$ in equation (\ref{eq:glines})
and in the finite disk factor $f$.
The lower limit assumes decelerating flows
($\partial v_{r} / \partial r < 0$) receive the same small force
contribution from nonradial rays as a flow that is not accelerating at all
($\partial v_{r} / \partial r =0$).
\subsection{Local Radiative Force Enhancement}
We induce azimuthal structure in our models by varying the Sobolev
line force over a localized ``star spot'' in the lower wind.
Since this force is directly proportional to the stellar flux
($L_{\ast} / 4 \pi r^{2}$), increasing or decreasing
$g_{r}^{\mbox{\footnotesize lines}}$ over a small area is
operationally equivalent to assuming
a bright or dark region of the photosphere.
Note, however, that by modulating the radiative force in this manner
we do {\em not} mean to literally propose the existence of
strong flux-varying spots on early-type stars.
We merely use this simple method to perturb the wind in lieu of more
definite knowledge about the physical cause(s) of surface
inhomogeneities.
Because the line driving grows weaker as one moves deeper into the
subsonic wind and photosphere, the force enhancement is essentially
confined to the transonic and supersonic wind, obviating the need
to model a perturbed stellar atmosphere.
The induced variation in the line force is assumed to have a specified
radial and azimuthal dependence which remains fixed to the stellar surface,
and thus rotates through the computational domain with the star.
The force enhancement is a function of radius $r$ and a
corotating azimuthal angle $\psi$,
\begin{equation}
\psi \, = \, \phi - \Omega t \,\,\,\, ,
\end{equation}
with $\Omega \equiv V_{\mbox{\footnotesize eq}}/ R_{\ast}$
the star's constant rotational angular velocity.
The perturbed line force has the form
\begin{eqnarray}
g_{r}^{\mbox{\footnotesize lines}} (r, \psi) &=&
g_{0} (r) + \delta g (r, \psi) \nonumber \\
&=& g_{0} (r) \left\{ 1 + A \,
\gamma (r) \exp \left[ - \left( \psi - \psi_{0} \right)^{2}
/ \sigma^{2} \right] \right\} \,\,\,\, ,
\label{eq:delg}
\end{eqnarray}
where $g_{0}$ is the unperturbed Sobolev
force (eq.~[\ref{eq:glines}]), $\psi_{0}$ is the azimuthal position
of the center of the spot, and $A$ is its dimensionless amplitude.
We will refer to ``dark spots'' as those with $-1 < A < 0$ and
``bright spots'' as those with $A>0$.
The azimuthal variation of the force is here assumed to be Gaussian
about $\psi_{0}$, with full width at half maximum (FWHM) given by
$\Phi \equiv 2 \sigma \sqrt{\ln 2}$.
The radial modulation $\gamma (r)$ is constrained by the geometrical
extent of the spot.
Close to the star, the spot is all that can be seen, and
so $\gamma (r \rightarrow R_{\ast}) \rightarrow 1$;
far from the star, the spot only represents a fraction
of the observed stellar disk, and
so $\gamma (r)$ approaches
a small, but constant value as $r \rightarrow \infty$.
For a field point directly over ($\psi = \psi_{0}$) a {\em circular}
flux enhancement with angular diameter $\Phi$,
the radial function $\gamma (r)$ can be derived analytically from the
the normalized residual flux,
\begin{equation}
\frac{{\cal F} - {\cal F}_{0}}{{\cal F}_{0}} \, = \,
A \frac{r^2}{R_{\ast}^{2}} \int \! \int D (r, \mu' , \phi' )
\, \mu' d \mu' d \phi' \, = \, A \gamma (r) \,\,\,\, ,
\end{equation}
where ${\cal F}$ and ${\cal F}_{0}$ are the total and unperturbed
fluxes.
The amplitude $A$ takes into account the relative magnitude of
the spot's ``residual effective temperature,'' and is equivalent to
$(T_{\mbox{\footnotesize spot}}^{4} - T_{0}^{4})/T_{0}^{4}$.
The area integral is taken over a solid angle centered about the
$z$-axis, with angles $\theta' = \cos^{-1} \mu'$ and $\phi'$ measured
from the field point in the wind at radius $r$, and the residual
limb darkening function $D$
set to zero for rays not intercepting the spot.
Thus, for simple linear limb darkening,
\begin{equation}
\gamma (r) \, = \, \frac{2 \pi r^{2}}{R_{\ast}^{2}}
\int_{\mu_{0}(r)}^{1} \frac{1}{4 \pi} \left( 2 + 3
\sqrt{ \frac{\mu'^{2} - \mu_{\ast}^{2}}{1 - \mu_{\ast}^{2}} }
\right) \mu' d \mu' \,\,\,\, ,
\label{eq:gamint}
\end{equation}
where $\mu_{\ast} \equiv [ 1 - (R_{\ast}^{2} / r^{2}) ]^{1/2}$
defines the stellar limb, and
\begin{equation}
\mu_{0} (r) \, = \, \left\{
\begin{array}{ll}
\mu_{\ast}, & r \cos (\Phi /2) \le R_{\ast} \\
\sqrt{1 - R_{\ast}^{2} \sin^{2} (\Phi /2) / S^{2}}, &
r \cos (\Phi /2) > R_{\ast}
\end{array} \right.
\end{equation}
defines the visible edge of the spot.
The distance $S$ from the field point to the edge of the spot is
$[ r^{2} + R_{\ast}^{2} - 2 r R_{\ast} \cos (\Phi /2) ]^{1/2}$.
We thus adopt the radial modulation function given by the analytic
integral of (\ref{eq:gamint}),
\begin{equation}
\gamma (r) \, = \, \frac{1}{2} \left[ 1 + \left(
\frac{1 - \mu_{0}^{2}}{1 - \mu_{\ast}^{2}} \right) -
\left( \frac{\mu_{0}^{2} - \mu_{\ast}^{2}}{1 - \mu_{\ast}^{2}}
\right)^{3/2} \right] \,\,\,\, ,
\label{eq:gamm}
\end{equation}
which approaches unity as $r \rightarrow R_{\ast}$ and approaches
a constant value of
$[1 + \sin^{2} (\Phi /2) - \cos^{3} (\Phi /2) ] /2$ as
$r \rightarrow \infty$.
Although this residual flux integral is also able to provide oblique
($\psi \neq \psi_{0}$) and nonradial components of the flux
enhancement of a star spot, we restrict our present models to the
explicit {\em radial} perturbation given by equations
(\ref{eq:delg}) and (\ref{eq:gamm}).
Again we emphasize that we do not mean to model in detail an
actual star spot, but are only using the spot-like force
enhancement as a convenient way to perturb the wind base.
Figure~1 shows contours of the force enhancement in the equatorial
plane for a spot with $\Phi = 20\arcdeg$,
as well as wind ``streaklines'' for various stellar rotation
speeds, as discussed further below (\S~3.3).
Note that the spot significantly affects only a relatively small
area of the wind: an azimuthal extent of $\sim 2 \Phi$ and a
radial extent of about a stellar radius.
This allows several spots to be superposed on the stellar surface
without any appreciable overlap in their force enhancements.
Following the empirical arguments of Kaper \& Henrichs \markcite{KH94}(1994),
who suggest a variable dipole magnetic field as the seed of large-scale
wind structure, we place {\em two} spots separated by $180\arcdeg$
on our model stars.
Before examining how a rotating wind responds to the localized force
enhancement, it is instructive to see how a non-rotating wind is
affected.
Figure~2 shows the radial velocity and density (directly over the spot)
at a reference radius of $10 R_{\ast}$, well beyond the region of
significant direct force enhancement, as a function of amplitude $A$.
The general trend is for a ``bright'' spot to increase the local
mass loss and thus increase the density of the wind near the star.
Further out in the wind, where the effect of
the spot drops off, the radiative force cannot accelerate the
higher density material as strongly,
so it approaches a {\em lower} terminal speed
than in the unperturbed wind.
Conversely, ``dark'' spots decrease the mass loss in the
surrounding wind and thus allow the less dense material to be
accelerated more strongly, leading to a much {\em higher} terminal speed.
The three sets of data in Figure~2 correspond to:
(1) one-dimensional, finite-disk, ``modified CAK'' (mCAK) solutions
with a realistic critical point analysis and numerical integration,
(2) two-dimensional hydrodynamical models using VH-1, and
(3) a simple analytic fit to the data.
The one-dimensional mCAK models contain the radial spot modulation
$\gamma (r)$, but no information about neighboring streamline divergence
or convergence.
For spots with $A \gtrsim 0.6$, the one-dimensional models cease to
have steady-state solutions that reach to infinity because too much mass
is driven off the star to be accelerated beyond its gravitational escape
velocity.
The two-dimensional models can drive more mass to infinity because
the density is reduced by a slight azimuthal expansion, which leads to
a faster-than-radial divergence of flow tubes (see, e.g., MacGregor
\markcite{M88}1988).
The simple fit to the velocity and density variation in Figure~2
depends on only one free parameter, and makes use of the approximate
dependence of the CAK mass loss rate on an arbitrary force multiplier
(see Cranmer \& Owocki \markcite{CO95}1995, eq.\ 24).
Directly over the spot,
\begin{equation}
\dot{M} \, \approx \, \dot{M}^{(0)} (1 + A)^{1/ \alpha} \,\,\, ,
\end{equation}
where $\dot{M}^{(0)}$ is the unperturbed mass loss rate, and we assume
that the CAK\markcite{CAK} critical point $r_{c}$, where the mass flux
is determined,
is close enough to the star that $\gamma (r_{c}) \approx 1$.
In a one-dimensional steady state, the mass conservation equation is
integrated in the usual way to obtain
$\dot{M} = 4\pi \rho v_{r} r^{2}$,
and this provides a relation between the velocity and density at a
given radius.
We thus define fitting functions which obey this multiplicative
constraint:
\begin{equation}
\rho \, \approx \, \rho^{(0)} (1+A)^{s/ \alpha} \,\,\, ,
\label{eq:nonrot}
\end{equation}
\begin{equation}
v_{r} \, \approx \, v_{r}^{(0)} (1+A)^{(1-s)/ \alpha} \,\,\, ,
\label{eq:nonrotv}
\end{equation}
where we have found the best fit value of $s=1.77$ for our dashed-line
fits in Figure~2.
\subsection{Numerical Specifications}
Let us next describe some of the details of our numerical discretization,
boundary conditions, and initial conditions.
We specify flow variables on a fixed two-dimensional spatial mesh
in radius $r$ and azimuthal angle $\phi$.
In our standard models we use 200 radial zones, from $R_{\ast}$
to $30 R_{\ast}$, with the zone spacing concentrated near the stellar
base where the flow gradients and spot enhancements are strongest.
The radial spacing starts at the lower boundary with
$\Delta r = 0.002 R_{\ast}$, then increases by 3\%
per zone out to a maximum of $\Delta r = 0.82 R_{\ast}$ at the
outer boundary.
The azimuthal mesh contains 160 constantly-spaced zones, ranging
from $0\arcdeg$ to $180\arcdeg$ with a spacing of 1$\fdg$125.
Limited test runs with double the resolution in radius and azimuth
showed some correspondingly greater detail in wind fine structure
(e.g., shocks and radiative-acoustic waves), but overall the results
were qualitatively similar to those for the standard resolution.
We specify the boundary conditions in our numerical method in two
phantom zones beyond each edge of the grid.
The azimuthal boundaries at $\phi = 0\arcdeg, 180\arcdeg$ are periodic,
ensuring symmetry in the full equatorial plane.
At the outer radial boundary, the wind is invariably supersonic
outward, and so we set the flow variables in the outer phantom zones
by a simple constant-gradient extrapolation.
The lower radial boundary of the wind is somewhat more problematic,
and we use the boundary conditions described by Owocki et
al.\ \markcite{OCB}(1994):
constant-slope extrapolation for $v_{r}$, rigid rotation for
$v_{\phi}$, and a fixed base density $\rho_{B}$.
Because the mass loss rates of line-driven winds are determined from
the equations of motion alone, we are able to specify an appropriate
``photospheric'' density
that yields a stable, subsonic boundary outflow
(see also Owocki, Castor, \& Rybicki \markcite{OCR}1988).
The time-dependent hydrodynamical
method requires a reasonable initial condition
to be specified over the entire grid at time $t=0$.
For this we relax an analytically derived mCAK model to a steady state
on a one-dimensional numerical grid, and then copy this onto the full
two-dimensional mesh.
This ensures that any time dependence results only from the
induced force perturbations.
The models are stepped forward in time at a fixed fraction (0.25)
of the standard Courant-Friedrichs-Lewy time step.
Because the radiative force enhancement, switched
on at $t=0$, only varies
in time by corotating with the stellar surface, the wind responds
by forming an outwardly moving ``front,'' behind
which the wind has settled to a rotating steady state.
The dynamical flow time for gas to radially cross the computational grid
is approximately $2 \times 10^{5} \, \mbox{s}$
for our unperturbed initial state wind.
Typically we find that models perturbed by star spots settle to a
corotating steady state within two
dynamical flow times, and so
we plot all models at
$t= 4 \times 10^{5} \, \mbox{s}$.
\section{Numerical Results}
Because the DAC phenomenon is primarily observed in O-star winds, we choose
to center our study on a standard model of
the O4f supergiant $\zeta$~Puppis.
Specifically, we take
$M_{\ast}=60 M_{\odot}$,
$R_{\ast}=19 R_{\odot}$,
$L_{\ast}=8 \times 10^{5} L_{\odot}$, and
$T_{\mbox{\footnotesize eff}} = 42,000 \, \mbox{K}$
(see, e.g., Howarth \& Prinja \markcite{HP89}1989;
Kudritzki et al.\ \markcite{Ke92}1992).
The measured rotational $V_{\mbox{\footnotesize eq}} \sin i$ for
$\zeta$~Puppis is $230 \, \mbox{km s}^{-1}$, which we take
for the equatorial rotation velocity of our standard model.
We neglect the small ($\sim$7\%) oblateness induced by this
degree of rotation, which corresponds to a Roche equipotential
surface rotating at 63\% of its critical angular velocity.
We assume an isothermal wind of temperature
$T_{\mbox{\footnotesize eff}}$, corresponding to a sound speed
$a= 24 \, \mbox{km s}^{-1}$, and use the line-driving constants
$\alpha = 0.60$ and $k=0.15$ (see eq.\ [\ref{eq:glines}]).
In one-dimensional mCAK models, these result in a terminal velocity
$v_{\infty} = 2580 \, \mbox{km s}^{-1}$ and a mass flux
$\dot{M} = 3.28 \times 10^{-6} \, M_{\odot} / \mbox{yr}$.
The base density for our subsonic lower boundary condition is
$\rho_{B} = 6 \times 10^{-11} \, \mbox{g cm}^{-3}$.
The localized ``star spot'' radiative force enhancement described above
depends primarily on two quantities: the amplitude $A$ and the
azimuthal full width $\Phi$.
These, together with the equatorial rotation velocity
$V_{\mbox{\footnotesize eq}}$, are the three free parameters we vary
in our study of non-axisymmetric structure formation.
Table~1 outlines the input parameters and several output quantities
(to be discussed below) for the models we computed.
Models~1 and 2 are standard ``bright'' and ``dark'' spot models,
with $A= +0.5$ and $-0.5$, and they represent a basis to explain the
general hydrodynamical phenomenon of stream interaction.
The subsequent models in Table~1 are intended to confirm our
understanding of the physics of CIR formation, and are
discussed below in a more limited fashion.
\subsection{Standard Bright Spot: Model 1}
Figure~3 shows gray-scale plots for the density, radial velocity,
azimuthal velocity, and radial Sobolev optical depth in Model~1,
{\em normalized to the unperturbed wind.}
To ease comparison with other models,
the azimuthal coordinates here have been incremented by a
constant factor to align the peak of the spot ($\psi_{0}$)
with the center-line or $x$-axis of the diagram.
Note the expected tendency for a bright spot to create higher
density and lower radial velocity.
However, the {\em azimuthal} velocity only differs by a small subsonic
amount from the unperturbed
angular-momentum-conserving form
$v_{\phi}^{(0)} (r) =
V_{\mbox{\footnotesize eq}} R_{\ast} /r$.
This demonstrates the almost purely {\em radial}
effect of the spot enhancement.
Most of the corotating structure from the spot settles onto nearly
constant spiral ``streaklines'' in the wind.
In the present models, streaklines are equivalent to flow streamlines in
the star's rotating frame of reference.
Figure~4 compares streamlines and streaklines computed for Model~1.
By numerically integrating the kinematic relation
\begin{equation}
\frac{r \, d\phi}{dr} \, = \, \frac{v_{\phi} (r,\phi) - r \Omega_{F}}
{v_{r}(r,\phi)} \,\,\, ,
\label{eq:streak}
\end{equation}
from a locus of points on the stellar surface
spaced evenly in $\phi$, one can alternatively compute either
streamlines in the inertial reference frame,
with $\Omega_{F} =0$, or
streaklines in the rotating reference frame,
with $\Omega_{F} =\Omega$.
Areas with a higher (lower) concentration of streaklines correspond
to regions of relative compression (rarefaction), though
not all density variations are reflected in the streaklines.
The dashed lines in Figure~4 show the streamline and streakline
originating directly over the spot.
The streamline appears nearly radial, but careful inspection
shows that it has a modest ($18\arcdeg$) prograde deflection,
resulting from corotation of the relatively slow wind outflow near
the surface.
Through most of the wind, however, the streamlines are
close to radial, and this allows one to qualitatively
interpret the azimuthal coordinate as a {\em time} dimension.
For any streamline at a fixed value of $\phi$, which is
intercepted by different streaklines as the (corotating steady
state) system sweeps by, the two-dimensional hydrodynamics becomes
effectively one-dimensional, but now truly time dependent.
This concept allows us to understand the ``spread out'' CIR structure
in terms of a simpler model of radial wave or shock propagation.
Hundhausen \markcite{H73}(1973) modeled solar-wind CIR formation and evolution
in one dimension using this approximation.
We can disentangle the actual patterns of high density, low velocity, and
more optically thick CIR structure by examining several
causally-connected regions of this model:
\begin{description}
\item[I.] {\bf Direct Enhancement:}
Close to the stellar surface, the Gaussian-shaped spot increases the
mass flux and wind density over a limited ($r \lesssim 2 R_{\ast}$)
region near the star.
This enhanced-density patch is slightly deformed by rotation from the
contours shown in Figure~1, but is essentially equivalent.
The density increases over the spot by a maximum factor of $\sim$2.6,
only slightly smaller than that predicted by the non-rotating analysis
(see eq.~[\ref{eq:nonrot}]).
This region also shows considerable azimuthal spreading in $v_{\phi}$
as the wind begins to adjust to the presence of the spot.
\item[II.] {\bf Prograde Precursor:}
Just ahead of the spot ($\psi > \psi_{0}$), a small fraction of the
enhanced higher-density wind is able to ``leak out'' azimuthally and
settle onto a set of relatively unperturbed streaklines.
The density in this feature is only enhanced by a factor of $\sim$1.2,
indicating that it comes from material in the prograde tail of the
Gaussian distribution.
It becomes isolated from the direct spot enhancement at a relatively
large distance from the star ($r \approx 6.4 R_{\ast}$)
after the CIR rarefaction (IV) has appeared
between it and the CIR shock (III).
\item[III.] {\bf CIR Compression:}
The low radial velocity wind from the center of the spot curls around on
more tightly-wound streaklines than the surrounding unperturbed wind, and
these streams begin to interact at a finite radius from the star
($r \approx 1.6 R_{\ast}$).
Alternately, in the above one-dimensional interpretation, the slow
stream can be considered equivalent to a radially-extended Gaussian
``wave packet'' which nonlinearly steepens as the fast mean wind
begins to overtake it.
The result of this collision of fast and slow streams is a corotating
weak shock compression (the CIR) which, because it is driven by ram
pressure from the mainly unperturbed
wind, propagates out at very near
the unperturbed wind velocity.
Because the flow is isothermal, we do not see a separation into a distinct
forward and reverse shock pair, as is observed in the more nearly adiabatic
solar wind.
\item[IV.] {\bf CIR Rarefaction:}
Ahead of the nonlinear shock the streaklines fan out and form a
lower-density rarefied region.
The formation of this rarefaction is mandated by mass flux
conservation, and the radial velocity correspondingly peaks slightly
above its unperturbed value here.
Because the density in this feature never dips too far below the
unperturbed density ($\min [\rho / \rho^{(0)}] \approx 0.82$),
the rarefaction propagates out at nearly the same velocity as the
CIR compression.
\item[V.] {\bf Radiative-Acoustic ``Kink:''}
In a purely hydrodynamical wind, the radial CIR shock structure is the
sole result of the nonlinear steepening of the initial enhancement.
Any nondissipative signals propagating in the rest frame
of the wind (at characteristic speeds $\pm a$) are limited to the
relatively undisturbed lateral (nonradial) direction.
In a line-driven wind, however,
Abbott \markcite{A80}(1980) and Rybicki et al.\ \markcite{Re90}(1990)
found that
large spatial-scale linear perturbations propagate in the radial
direction at modified ``radiative-acoustic'' characteristic speeds.
Abbott \markcite{A80}(1980) derived
\begin{equation}
C_{\pm} \, = \, -\case{1}{2} U \pm \sqrt{ \left( \case{1}{2} U
\right)^{2} + a^{2} }
\end{equation}
for radial modes, where $U \equiv \partial (
g_{r}^{\mbox{\footnotesize lines}} ) / \partial ( \partial v_{r} /
\partial r )$ and $C_{\pm}$ reduces to the purely acoustic case if
$U=0$.
In most of the wind, though, $U \gg a$, and the outward (positive root)
solutions are subsonic, and the inward (negative root) solutions are
supersonic.
In Model~1 we see both an acoustic lateral mode (spreading in
$v_{\phi}$ at large radii; Figure~3c) and a nonlinear analog of the
inward radiative-acoustic mode, which propagates
slowly outward in the star's frame ($0 < v_{r} + C_{-} < v_{r}$)
as a weak discontinuity, or ``kink,'' in the
radial velocity gradient.
Because of its slow propagation (more tightly-wound streaklines)
this feature eventually collides with the CIR rarefaction from the
other spot at a radius of $\sim$13.5~$R_{\ast}$ and ceases to exist.
\end{description}
We trace these five features
in Figure~5, which is a
close-up of the density gray-scale shown in Figure~3a.
Unique tracks were found by searching for local extrema (in radius)
of various quantities, and following contiguous patterns around in
azimuth.
The direct spot enhancement (I) appears at the stellar surface
as a local maximum in the
normalized density $\rho / \rho^{(0)}$, and collides with the
CIR/kink pair of features (III and V) at a radius of
$\sim$3.9$R_{\ast}$.
These ``bifurcated'' extrema are found by tracking local minima and
maxima in the radial velocity, as shown in Figure~6 below.
They appear at a relatively small radius $r_{L}$ (see Table~1),
where the spot perturbation is still linear.
The remaining precursor/rarefaction pair of features (II and IV)
correspond to other local minima and maxima in the normalized density,
and they appear further out (at a larger radius $r_{NL}$) where
the disturbance has definitely steepened into a nonlinear shock.
It is interesting that the CIR compression and rarefaction do not
form together at the same point, but this is understandable, since the
latter can be considered an effect or response of the former.
To get an indication of which wind structures should yield the most
prominent signatures in observed line profile variations, let us examine
the radial Sobolev optical depth,
\begin{equation}
\tau_{r} (r,\phi) \, \equiv \, \frac{\kappa_{L} \, \rho (r,\phi)
\, v_{\mbox{\footnotesize th}}}{\left| \partial v_{r} (r,\phi) /
\partial r \right|} \,\,\, .
\label{eq:tausob}
\end{equation}
The gray-scale plot in Figure~3d shows the changes in the optical depth
relative to the mean, unperturbed wind.
Since both the line absorption coefficient $\kappa_{L}$ and
the ion thermal speed $v_{\mbox{\footnotesize th}}$ are assumed to have
the same constant values in both the mean and perturbed flow, all variations
here stem from changes in the ratio of density to velocity gradient.
Somewhat surprisingly, the regions of strongest optical depth enhancement
occur not within the dense CIR compression (feature III), but rather within
the relatively shallow velocity gradient region after the Abbott kink
(feature V).
Figure~6 plots the radial variation of velocity and density from selected
slices of constant azimuthal angle $\phi$.
For this corotating steady state, the changing features in these line plots
also indicate the time evolution of structure at fixed azimuths.
This allows us to follow the outward propagation of both
the CIR density enhancement (local minima in velocity) and the
trailing radiative-acoustic kink (local maxima in velocity).
In the high-density CIR, the wind is either strongly accelerating or
decelerating, so both the numerator and the denominator in the Sobolev optical
depth (eq.~[\ref{eq:tausob}]) are enhanced, resulting in very little
net increase.
Just outward from the kink, however, the density is nearly unperturbed,
while the velocity gradient is much shallower, implying a large increase
in the Sobolev optical depth.
In synthetic line profiles (see \S~4.2 and Figure~11a), this near
plateau produces a distinct absorption feature quite similar to
slowly evolving DACs.
As the (corotating) steady-state structure rotates in front of the
observer's line of sight, material flowing {\em through} the kink appears
at the velocities of the local maxima in Figure~6,
but the evolution of the feature is governed by the
radiative-acoustic mode propagation, which leads to an
apparent acceleration that is much slower than the
actual acceleration of the wind material.
Ahead of the CIR, between the compressive
density maximum and the rarefied minimum, there is a region
of high acceleration that arises from the
prograde edge ($\psi > \psi_{0}$) of the Gaussian spot, which
steepens into a monotonic sawtooth structure connecting
the shock with the lower-density unperturbed wind.
This region contains a lower net optical depth than the unperturbed
wind, implying a relative {\em lack} of absorption in synthetic line
profiles.
This effect is only slightly weaker than the enhanced absorption due
to the plateau, suggesting that isolated patches of extra
absorption (DACs) may be difficult to model theoretically without a
corresponding lower optical depth feature (with apparent relative
``emission'').
\subsection{Standard Dark Spot: Model 2}
A wind perturbed by a locally decreased radiative force
($A = -0.5$) produces a lower-density, high-speed stream, and thus
settles to a steady state on the computational grid faster than
a model with slow streams.
Figure~7 shows gray-scale plots for the density, radial velocity,
azimuthal velocity, and radial Sobolev optical depth for Model~2,
normalized in the same way as in Figure~3.
Although the dark spot produces an extremely rarefied wind, a
high-density CIR forms (on the leading edge of the perturbation) where
the high velocity stream collides with the slower unperturbed wind.
The corotating structure present in Model~2 is qualitatively simpler
than that in Model~1.
The CIR/kink pair of features initially appears at
$r_{L} \approx 2.2 R_{\ast}$, and advects smoothly throughout the wind.
There is no analog to the second pair of features (starting further
out at $r_{NL}$) in this model.
For slices of constant $\phi$, Figure~8 plots the radial dependence
of the velocity and density.
The contrast with the slow structure in
Figure~6 is apparent.
Note that the back-propagating radiative-acoustic kink is also
present in this model, comprising the left edge of the flat-topped
velocity peaks.
However, since the kink here is a reaction to the forward-steepened
shock, it forms within the fast and rarefied upstream region, and thus
does not contribute strongly to the Sobolev optical depth of the
corotating feature.
The high density CIR also does not have an enhanced opacity
because of the steep velocity gradients near the shock, as in Model~1.
Indeed, Figure~7d shows that most of the highest optical depth
material comes from the unperturbed wind, and that the high-speed
CIR should be mainly a source of {\em decreased} absorption.
The isolated ``clumps'' of highest optical depth in Figure~7d
(and in Figure~3d) are artifacts of the finite differencing used
to compute the radial velocity gradient, and do not significantly
affect the volume-integrated quantities used in constructing
line profile diagnostics.
The CIR structure in Model~2 appears very similar to that seen in
hydrodynamic models of the solar wind, e.g., note the resemblance
between Figure~8 and Figures~2 and 3 of Hundhausen \markcite{H73}(1973).
The two major differences between solar wind high-speed streams and
those in our model are:
(1) the distinct forward and reverse adiabatic shocks in the former and
(2) the back-propagating radiative-acoustic kink in the latter.
Despite this similarity, all subsequent models in our parameter study
use the bright spot of Model~1, which promises to simulate better the
slow DACs in early-type stellar winds.
\subsection{Variation of Spot Amplitude, Width, and Stellar
Rotation Velocity}
Let us now examine the effect of changing various spot and wind parameters.
To provide a basis for understanding the full hydrodynamical calculations,
we can estimate the expected effects in terms of a simple wind
streakline picture (see Figure~4).
Note that the shape of these streaklines can be
well approximated by neglecting both the wind's acceleration and
the angular-momentum-conserving azimuthal velocity, which nearly ``cancel
each other out'' when computing streakline deflection. Thus, the
angular deflection is given in this approximation by the Archimedean
spiral relation,
\begin{equation}
\phi - \phi_{0} \, \approx \, - \frac{\Omega}{v_{\infty}} \,
(r - r_{0}) \,\,\,\, .
\label{eq:archimedes}
\end{equation}
Mullan \markcite{M84a}(1984a)
used this to estimate the interaction radius $r_{i}$ between
a fast and slow stream initially set apart on the stellar surface by
a given azimuthal
separation $\Delta \phi$,
\begin{equation}
\frac{r_{i}}{R_{\ast}} \, = \, 1 +
\frac{\Delta \phi}{V_{\mbox{\footnotesize eq}}} \left(
\frac{v_{\infty}^{(f)} v_{\infty}^{(s)}}
{v_{\infty}^{(f)} - v_{\infty}^{(s)}} \right) \,\,\, .
\label{eq:ri}
\end{equation}
Here
$v_{\infty}^{(f)}$ and $v_{\infty}^{(s)}$ are the terminal velocities
of the fast and slow streams.
Table~1 contains this simple prediction for $r_{i}$ for all the
hydrodynamical models in our parameter study.
We assume $\Delta \phi = \Phi$, and for bright-spot models, we
take the fast stream to be the unperturbed wind, and the slow stream
to have $v_{\infty}^{(s)}$ given by equation~(\ref{eq:nonrotv}).
Conversely, for the dark-spot model the fast stream is given by
equation~(\ref{eq:nonrotv}) and the slow stream is unperturbed wind.
Models~3A, 3B, and 3C vary the spot amplitude $A$, and thus the
direct enhancement in density and velocity over the spot (see Table~1).
The resulting structure looks qualitatively similar to that of Model~1,
but the fast/slow stream interaction takes place at different radii.
Figure~9a traces the corresponding CIR compression and
radiative-acoustic kink features for these models.
As expected, CIRs form further out when there is a
smaller discrepancy between the fast and slow stream speeds,
but the linear minima and maxima in velocity first appear
(at $r_{L}$) much closer to the star than the simple interaction
analysis above predicts.
The ``nonlinear'' radius $r_{NL}$, where the CIR rarefaction
branches off from the prograde dense precursor, is relatively
{\em constant} with $A$, indicating a qualitatively different
formation mechanism from the compression/kink features.
Figure~10 shows how the velocity law of the kink and CIR shock varies
with spot amplitude, with each model rotated in azimuth to line up
similar features.
For the smallest values of $A$ (0.01, 0.1), no shock has yet formed,
so both features appear ``kink-like,'' propagating nearer
to the slow radiative-acoustic mode speed than to the mean wind speed
(see Figure~9a).
The shock steepening is evident for larger values of $A$, and the most
extreme model ($A=2.50$) shows considerable slowing of the dense
CIR resulting from the inverse dependence of the radiative force with
density (eq.~[\ref{eq:glines}]).
Our canonical Model~1 CIR, then, seems just on the verge between both
``slowing'' mechanisms, as well has having a shape between the
low-$A$ weak discontinuities and the high-$A$ shocks.
In all cases, however, the decelerating plateau-region ahead of the
kink has the same characteristic (negative) acceleration, indicating
that the enhanced optical depth at this radius may not vary strongly
with spot amplitude.
Models~4A, 4B, and 4C vary the azimuthal spot width $\Phi$, and Figure~9b
shows their CIR compression and kink features.
Increasing or decreasing the full-width $\Phi$ simply alters the
spatial scale of the interactions, and, in the limit of very small
spots (where the star's sphericity can be neglected) the CIR structures
seem self-similar with respect to an overall expansion factor.
Table~1 indicates an inverse relationship between $\Phi$ and the
maximum CIR density, and this can be understood heuristically by the
fact that when a given spot amplitude is spread over a larger area,
the collision of fast and slow streams is more diluted, and the
resulting shocks are not as strong.
Model~4A, with the smallest width $\Phi=10\arcdeg$, shows an apparent
reversal in this trend, and we suspect that some of the detailed
shock structure is under-sampled in our relatively coarse (in this
case) azimuthal grid.
Note that Mullan's \markcite{M84a}(1984a)
interaction radius $r_{i}$ agrees well with
the location of the nonlinear rarefaction/precursor feature $r_{NL}$
for these models.
Observations (Prinja \markcite{P88}1988; Henrichs et
al.\ \markcite{HKZ}1988)
suggest that the recurrence and acceleration times of DACs tend to
vary inversely with the projected equatorial rotation speed,
$V_{\mbox{\footnotesize eq}} \sin i$.
To examine how well our dynamical models might reproduce these observed
trends, Models~5A and 5B vary the equatorial stellar rotation velocity
$V_{\mbox{\footnotesize eq}}$.
Figure~9c shows their CIR compression and kink features.
The dominant effects are the overall variation of the spiral streaklines
(see, e.g., eq.~[\ref{eq:archimedes}]), and the
inverse centrifugal dependence of
$v_{\infty}$ on the rotation rate (Friend \& Abbott \markcite{FA86}1986).
The strength of the CIR shock decreases for more slowly rotating stars,
and we believe that this results from the variation of the streaklines
with respect to the star-spot enhancement (Figure~1).
Perturbed gas which rapidly advects out of the region of direct enhancement
receives less of a ``boost'' of extra density, and thus does not form
as strong a compression when interacting with the ambient wind.
\section{Synthetic Observational Diagnostics}
\subsection{SEI Line Profile Construction}
Ultraviolet P~Cygni lines are sensitive probes of the wind structure
of hot luminous stars, and it is important to model accurately
observed variations in their profile shape.
Here we use a multidimensional extension of the SEI (Sobolev with Exact
Integration) method of Lamers, Cerruti-Sola, \& Perinotto
\markcite{Le87}(1987) to
compute synthetic line profiles.
In this method, the source function is calculated locally in the wind
using the Sobolev escape probability approximation, and the emergent
flux profile is computed by numerically integrating the formal solution
to the transfer equation.
Bjorkman et al.\ \markcite{Be94}(1994)
discuss a two-dimensional extension of the
basic SEI algorithm, and our computational approach is similar in that
we do not yet treat the general case of nonlocal (or line doublet)
resonance coupling.
Our method, however, efficiently computes line profiles from an
arbitrary three-dimensional distribution of density and velocity,
for observers at arbitrary vantage points.
Following the notation of POF\markcite{POF}, we parameterize the
opacity of a model pure-scattering resonance line by defining
a dimensionless line-strength
\begin{equation}
k_{L} \, \equiv \, \left(
\frac{\dot{M} v_{\mbox{\footnotesize th}}}
{4 \pi R_{\ast} v_{\infty}^{2}}
\right) \kappa_{L} \,\,\, ,
\end{equation}
where $v_{\infty}$ and $\dot{M}$ are taken from the unperturbed model wind.
We assume the mass absorption coefficient
$\kappa_{L}$ is constant in radius, which is valid for
lines of interest in the dominant ionization stage of the wind, and at
least allows qualitative comparison for other lines.
We consider two representative cases: a moderate unsaturated line
($k_{L}=1$) and a strong saturated line ($k_{L}=100$).
The local two-dimensional (solid angle) integrals required to obtain the
escape probability and core-penetration probability for the Sobolev
source function are computed using Romberg's successive refinement
algorithm.
The explicit form of these integrals is found in Lamers et
al.\ \markcite{Le87}(1987).
POF\markcite{POF} compare the use of the local Sobolev formalism
with a self-consistent multiple-resonance technique in structured wind
models, and find significant disagreement in the resulting line
profiles.
However, our CIR model winds are much less structured than the
one-dimensional instability models used by POF\markcite{POF},
with at most only two zones of nonmonotonic velocity
variation in the entire wind.
Further, here we concentrate on {\em residual} line profile
variability, which should be
less susceptible to consistency errors in the source function than
the actual line profile shape.
We perform the ``exact integration'' for the line flux using a
cylindrical $(p', \phi', z')$ coordinate system with the observer
oriented along the positive $z'$-axis at an infinite distance from
the origin.
The equation of radiative transfer is evaluated in differential form
along rays parallel to this axis, and along each ray the specific
intensity is integrated using second order implicit Euler differencing.
The resulting emergent intensities at the outer boundary of the
computational grid are then integrated by nested Romberg quadrature
in $p'$ and $\phi'$ to form the flux, and this process is repeated
for each frequency point in the total line profile.
We refer the reader to Lamers et al.\ \markcite{Le87}(1987) and
POF\markcite{POF}, who summarize
these integrals, but only apply them in the spherically symmetric
($\phi'$ independent) case.
We make two major approximations in our line profile construction
technique:
\begin{enumerate}
\item The three-dimensional wind structure is formed by interpolating
in latitude between the two-dimensional equatorial plane models and
the one-dimensional unperturbed polar wind.
For simplicity we assume the same Gaussian structure of the ``star
spots'' in latitude as in longitude, and apply it to the entire wind:
\begin{eqnarray}
\rho (r,\theta,\phi) &=& \rho^{(0)} (r) \, [1 - E(\theta)] \, + \,
\rho^{\mbox{\footnotesize (2D)}}(r,\phi) \, E(\theta) \nonumber \\
v_{r} (r,\theta,\phi) &=& v_{r}^{(0)} (r) \, [1 - E(\theta)] \, + \,
v_{r}^{\mbox{\footnotesize (2D)}}(r,\phi) \, E(\theta) \\
v_{\phi} (r,\theta,\phi) &=& v_{\phi}^{(0)} (r) \,
[1 - E(\theta)] \, \sin\theta \, + \,
v_{r}^{\mbox{\footnotesize (2D)}}(r,\phi) \, E(\theta)
\, \sin\theta \,\,\,\, , \nonumber
\end{eqnarray}
where $E(\theta) \equiv \exp [ -( \pi/2 - \theta)^{2} / \sigma^{2} ]$
(see eq.~[\ref{eq:delg}]).
As in the two-dimensional models, we retain our assumption that
$v_{\theta}=0$.
Note that the azimuthal velocity has an extra factor of $\sin\theta$
included to preserve angular momentum conservation out of the equatorial
plane, and this provides a latitudinal wind variation even for
{\em unperturbed} models.
\item
The total Doppler width of the Gaussian line profile contains both
thermal and microturbulent contributions,
\begin{equation}
v_{D} \, \equiv \, \sqrt{
v_{\mbox{\footnotesize th}}^{2} +
v_{\mbox{\footnotesize turb}}^{2} } \,\,\,\, ,
\end{equation}
with the thermal speed set by $v_{\mbox{\footnotesize th}} = 0.28 a$,
as appropriate for CNO driving ions.
For simplicity, we assume a constant microturbulent velocity
$v_{\mbox{\footnotesize turb}}= 100 \, \mbox{km s}^{-1}$, though
better line-profile fits have been obtained by assuming this
varies in proportion to the mean wind velocity
(Haser et al.\ \markcite{He95}1995).
Though phenomenological, this use of a microturbulent velocity allows
realistic line-profile synthesis with a minimum of free parameters.
It also compensates for the suppression here of the small-scale instability,
which one-dimensional, nonlocal simulations have shown to result in
many of the same observational signatures as microturbulence
(POF\markcite{POF}; Owocki \markcite{O94}1994).
\end{enumerate}
\subsection{Time Variability in Dynamical Models}
We produce time-variable P~Cygni line spectra by positioning an
``observer'' in the equatorial plane
at successive azimuthal angles $\phi$ with respect to the
two-dimensional models.
For the standard rotation velocity used in Models 1, 2, 3, and 4,
the line profile variability repeats with a period
\begin{equation}
\Pi \, = \, \frac{\pi R_{\ast}}{V_{\mbox{\footnotesize eq}}}
\, = \, 2.091 \, \mbox{days} \,\,\, .
\end{equation}
Of course, since spiral streakline structures in the models often
subtend more then 180$\arcdeg \,$ of azimuth, continuous DAC-like signatures
can exist for times longer than this period.
We arbitrarily define $t=0$ when the observer is positioned directly over
the center of one of the photospheric spot perturbations.
As emphasized by Lamers \markcite{L94}(1994),
we find that most of the line profile
variability occurs in the absorption column of the wind
($0 \leq p' \leq R_{\ast}$),
with comparatively little variation in the
larger emission volume ($p' > R_{\ast}$).
Because the flux integration over this emission
volume dominates the CPU time in our SEI line-profile synthesis,
we only computed full, variable emission-volume profiles for
a few selected test cases, namely for
observers at four equally-spaced azimuthal
angles ($\phi = 0\arcdeg, 45\arcdeg, 90\arcdeg, 135\arcdeg$) for Model~1,
as well as at one arbitrary azimuthal angle for an unperturbed
$\phi$-independent wind model.
For an unsaturated ($k_{L} = 1$) line,
we found that the perturbations in the absorption-column flux
reach as high as 47\% of the continuum level, whereas those in
emission-volume flux never exceed 1.9\%.
Since these latter variations would only be
marginally observable with IUE signal-to-noise
ratios of 20 to 40, we neglect the perturbed emission
volume when computing subsequent line profiles on a finer time-
and velocity-sampled grid.
Figures~11a and 11b show the absorption-column line profile variability
for Model~1, repeated over three data periods
(1.5 rotation periods) to emphasize the
rotationally-modulated structure.
Following the standard observational convention, the gray-scale is
normalized by a ``minimum absorption'' (maximum flux)
template, constructed independently at each line velocity.
This choice contains the implicit bias that the variability
takes the form of extra {\em absorption,} which is
only partially appropriate for Model~1.
The unsaturated ($k_{L}=1$) line exhibits definite DACs that apparently
accelerate through the profile on $\sim$3.9-day time scales, even
though their recurrence time is shorter.
The saturated ($k_{L}=100$) line exhibits blue-edge variability
on the same time scale.
Figures~11a and 11b also contain the
time-averaged line profiles for Model~1 and the
minimum and maximum absorption templates, which show the extent
of the absorption variability at each velocity.
We also plot the standard deviation of the data, allowing
a qualitative comparison to observed temporal variance spectra.
Figure~12 shows the absorption-column line profile variability for
Model~2.
The accelerating features in this unsaturated ($k_{L}=1$) line
differ from those in Figure~11a in two apparent ways.
First, because of the high-velocity stream induced by the dark spot,
the variability extends out to nearly 5000 km~s$^{-1}$, almost twice
the unperturbed wind terminal speed.
Second, the enhanced absorption at lower line velocities
($v \lesssim 2000 \, \mbox{km s}^{-1}$) represents the mean state,
and the isolated accelerating features appear as a lack of absorption.
Although we anticipated this trend in \S~3.2, it is surprising to
note the overall {\em similarity} between the bright and dark spot
profiles at higher line velocities
($2000 \, \mbox{km s}^{-1} \lesssim v \lesssim 2400 \, \mbox{km s}^{-1}$).
The primary difference here is that the strongest absorption feature trails,
in Model~1, and leads, in Model~2, the overall region of enhanced
absorption.
Far from the star, the near-terminal-speed wind is beginning to
laterally homogenize the bright or dark spot perturbations into a
simpler pattern of alternating large-scale compressions and rarefactions.
In order to interpret more clearly the evolution of DACs in our models,
we utilize the observational fitting technique of Henrichs et
al.\ \markcite{He83}(1983) and Kaper \markcite{K93}(1993),
which represents the DACs as
dense, plane-parallel slabs of gas in the observer's line of sight.
For each component, the dependence of the quotient flux (normalized
by the minimum absorption template) on the
line-of-sight velocity $v$ is fit by
\begin{equation}
I(v) \, = \, \exp \left\{ - \tau_{c} \exp \left[ - \left(
\frac{v - v_{c}}{v_t} \right)^{2} \right] \right\} \,\,\, ,
\end{equation}
where $\tau_{c}$ is a representative central optical depth,
$v_{c}$ is the line velocity of the center of the DAC, and $v_{t}$
is related to its width in velocity space.
These three parameters are varied and fit to each feature in the
time series using Marquardt's $\chi^2$ method
(Bevington \markcite{B69}1969).
Although many of our synthetic DACs are asymmetric about $v_{c}$,
with slightly more absorption on the low-velocity side of the feature,
the fits always reproduce well the overall line shape.
One additional useful quantity, the column density of the DAC, is
given for our model lines by
\begin{equation}
N_{col} \, = \, \frac{\sqrt{\pi}}
{\kappa_{L} v_{\mbox{\footnotesize th}} \langle m \rangle}
\frac{\tau_{c} v_{t}}{(1 + v_{c}/c)} \,\,\, ,
\end{equation}
where $\langle m \rangle$ is the mean mass of gas atoms and ions
(see POF\markcite{POF} and Kaper \markcite{K93}1993).
In Figure~13 we plot the resulting fit parameters
($v_{c}$, $v_{t}$, $N_{col}$) as a function of
time for the bright-spot, unsaturated-line models in the parameter study.
The inherent ``overlap'' in the time series (i.e., multiple DACs at
a given time) has been removed to more clearly show the evolution
of the individual DAC.
As is often seen in observed line profile variability, the feature
accelerates through the line profile while growing progressively
narrower, its column density increasing to a maximum value, then
decreasing as the DAC nears its terminal velocity.
In our models, $v_{c}$ often reaches or exceeds the wind's unperturbed
$v_{\infty}$, but the rapidly dropping values of
$N_{col}$ (which also is related to the equivalent
width of the DAC) might preclude actual observation of this final period
of evolution.
In fact, most of the DACs we track seem to approach an initial
``pseudo terminal speed" ($\sim$~0.8-0.9~$v_{\infty}$) while the column
density is at its peak, then accelerate further to the wind's terminal
speed as the column density decreases.
As expected, the DACs produced by radiative-acoustic kinks accelerate
quite slowly.
Figure~13c compares the acceleration of DACs from Models 1, 5A, and
5B with several analytic ``beta'' velocity laws.
We compute $v(t)$ by numerically integrating the kinematic relation
\begin{equation}
t \, = \, t (r [v]) \, = \,
\int_{R_{\ast}}^{r} \frac{dr'}{v_{0} + (v_{\infty}-v_{0})
(1 - R_{\ast}/r')^{\beta}} \,\,\, ,
\end{equation}
where we take $v_{0} = a$ and $v_{\infty} = 2580 \, \mbox{km s}^{-1}$.
The slow acceleration of the DACs is equivalent to
$\beta \approx 2-4$, which agrees with the observations of
Prinja et al.\ \markcite{Pe92}(1992) and
Prinja \markcite{P94}(1994).
Note, however, that an estimation of a single characteristic $\beta$
for a DAC is problematic, since
(1) its terminal speed is not clearly defined, and
(2) $v_{c}(t)$ experiences several minor acceleration and deceleration
episodes superimposed on the overall DAC acceleration.
Figure~14 plots acceleration versus velocity for the DAC of Model~1,
and compares it to the beta laws defined above.
Note the similarity in both magnitude and nonmonotonic behavior between
this theoretical acceleration and that found by Prinja
et al.\ \markcite{Pe92}(1992) for DACs in the wind of $\zeta$~Puppis.
\section{Summary, Conclusions, and Future Work}
We have carried out two-dimensional hydrodynamical simulations of an
azimuthally inhomogeneous radiation-driven wind from a rotating O star.
The wind responds to a photospheric radiative force enhancement by
forming large-scale corotating structures extending far beyond the
region of direct perturbation.
Although classical CIR compressions and rarefactions are often seen,
the most important structures observationally are slowly-propagating
radiative-acoustic kinks or plateaus which have a large Sobolev
optical depth.
These plateaus show up as strong DACs when formed behind streams
resulting from {\em enhanced
mass loss,} and they accelerate at a slower rate than the wind
passes through them.
Preliminary P~Cygni line profile synthesis has shown several important
trends in the emergent DACs from these structures:
\begin{enumerate}
\item The slow acceleration of DAC features is fit reasonably well by
$\beta \approx 2-4$, and we find no significant correlation between
their acceleration time scale and the star's rotation velocity.
Of course, since the CIRs in our models are linked to the rotating
surface, there is a definite correlation between the {\em recurrence}
time scale of DACs and $V_{\mbox{\footnotesize eq}}$.
\item Despite minor variations, the DAC parameters $v_{c}$ and $v_{t}$
do not depend strongly on the amplitude $A$ or full width $\Phi$ of
the model spot perturbation.
The primary exception is the high-amplitude Model~3C ($A=2.50$) which
accelerates much more slowly due to the higher density in the
CIR shock.
\item The optical depth $\tau_{c}$ and column density $N_{col}$
seem to be sensitive probes of the amplitude of the initial
surface perturbation (see Figure~13a, bottom panel).
The concavity of $N_{col}$ during the strongest period of DAC
evolution may be able to provide information about the azimuthal
size of the perturbation, but this is not
as clearly observable an effect
(see Figure~13b, bottom panel).
\end{enumerate}
Let us now compare our model results with actual observations of large-scale
wind structure.
In many cases both (1) slow and quasi-episodic DACs and (2) faster
periodic modulations are seen simultaneously in OB-star winds
(Kaper \markcite{K93}1993; Massa et al.\ \markcite{Me95}1995).
As seen above, the first type of variation can be readily reproduced
by the CIR and radiative-acoustic plateau that results from an
azimuthally localized mass-loss enhancement.
Though our models assume strict rotational periodicity,
such DACs could also be readily caused by {\em transient} CIRs
for which the surface mass loss ``eruption'' lasts long enough
to make structure that covers a substantial portion of the stellar disk.
The second type of structure observed in OB winds (fast,
near-sinusoidal flux variations accelerating {\em with} the wind) is
more difficult to produce with a CIR model,
because the required optical depth or density variations disrupt
the mean wind velocity streaklines.
To explain these faster modulations, we intend to investigate models in
which wind structure is induced by nonradial pulsation (NRP) of
the underlying star
(Owocki, Cranmer, \& Fullerton \markcite{Oe95}1995).
The success here in reproducing realistic DACs suggests that the
CIR model warrants further study and development.
One important extension will be to use a more complete radiative force
that incorporates the line-driven instability.
The resulting stochastic variations may disrupt the large-scale
CIR structure, but slowly evolving kink-like plateaus
have been seen to survive and propagate
in various one-dimensional instability simulations
(Fullerton \& Owocki \markcite{FO92}1992;
Owocki, Fullerton, and Puls \markcite{Oe94}1994).
In addition, extending the present models to three dimensions
may shed light on the variability of wind-compressed disks and other
latitudinally-varying structures in, e.g., Be-star winds.
Finally, it will be important
to develop techniques to synthesize
other observational diagnostics, such as subordinate-level UV
and optical lines, infrared photometry, and continuum and line
polarization, which will allow further constraints
on models of wind variability.
The resulting phenomenological ``atlas'' would provide a solid basis
for interpreting the great diversity of OB-star wind variability
observations in terms of fundamental, dynamical models of
time-dependent wind structure.
\acknowledgments
This work was supported in part by
NSF grant AST 91-15136 and
NASA grant NAGW-2624 (to SPO),
and a NASA Graduate Student Researcher's Program fellowship (to SRC).
Supporting computations were made possible by an allocation from
the San Diego Supercomputer Center.
We thank A.\ Fullerton, K.\ Gayley, D.\ Massa, and D.\ Mullan
for many helpful discussions and comments.
We also thank J.\ Blondin for initially providing the VH-1
hydrodynamics code.
\newpage
|
\section{Introduction}
The effective action has turned out to be a quite important
subject in
the study of different aspects of quantum field theory. Among
the
phenomena to which it has been applied successfully, we can
mention
symmetry breaking/restoration effects, phase transitions in
general,
models of quantum corrected field equations, etc.
Most of the studies of the effective action have been limited to
a
quasi-local approach (for a general introduction see \cite{8,1}),
that
is, they deal with almost constant background fields, as is the
case of
quantum gravity on a De Sitter background \cite{18} ---which is
important in inflationary universes. Recently, some interest has
arisen
(\cite{9}-\cite{12},
see \cite{11} for an extense account) for the case of weak but
very
quickly varying background fields, which typically lead to
non-local
effective actions. In the present note, by using simple
renormalization
group (RG) methods ---implemented by means of a Wilsonian
procedure \cite{5}--- we are going to show how one can obtain in
fact an
improved non-local effective gravitational action for a big class
of
theories.
The starting point for our considerations will be a massless,
multiplicatively renormalizable theory including scalar, spinor
and
vector fields on a classical gravitational background. The
corresponding Euclidean Lagrangian has the following form
\begin{eqnarray}
L &=& L_m + L_{ext}, \nonumber \\
L_m &=& L_{YM} + \frac{1}{2}
(\nabla_\mu \varphi )^2 +
\frac{1}{2} \xi R \varphi^2
+ \frac{1}{4!} f \varphi^4
+i\overline{\psi} (\gamma^\mu \nabla_\mu-h \varphi ) \psi, \nonumber \\
L_{ext} &=& a_1R^2 + a_2 C^2_{\mu\nu\alpha\beta} +a_3 G + a_4
\Box R.
\label{541}
\end{eqnarray}
By choosing a specific gauge group, we can assume that some
multiplets
of the scalar, $\varphi$, and spinor, $\psi$, fields are given in
some
concrete representation of the gauge group.
We will assume that our theory (\ref{541}) is a typical
asymptotically
free GUT in curved spacetime (for a general introduction, see
\cite{1}).
In principle one could equally well consider other types of GUTs,
what
would not change qualitatively the conclusions of our study
below.
The running coupling constants corresponding to the
asymptotically free
couplings of the theory (\ref{541}) have the form \cite{3,4}
\begin{eqnarray}
&& g^2(t) =g^2 \left[ 1 + \frac{B^2g^2t}{(4\pi)^2} \right]^{-1},
\ \ \ \ g^2(0)=g^2, \nonumber \\
&& h^2(t) =\kappa_1g^2(t) , \ \ \ \ \ \
f(t) =\kappa_2g^2(t),
\label{542}
\end{eqnarray}
where $t$ is the RG parameter while $\kappa_1$ and $\kappa_2$
are
numerical couplings
defined by the specific features of the theory under
consideration. We
know of many examples of such theories, with gauge groups SU(N),
O(N),
E$_6$, etc. \cite{3,4}.
Asymptotic freedom ($g^2(t) \rightarrow 0$, $t \rightarrow
\infty$) is
realized for all running couplings: gauge,
Yukawa and scalar ones, as is easy to see from (\ref{542}).
The study of asymptotically free GUTs in curved
spacetime was started in Ref. \cite{2} (for a review and detailed
list
of references see \cite{1}). In the theories with one scalar
multiplet,
for the running scalar-graviton coupling constant one gets
\begin{equation}
\xi (t) =\frac{1}{6} + \left( \xi - \frac{1}{6} \right) \left[ 1
+
\frac{B^2g^2t}{(4\pi)^2} \right]^b,
\label{543}
\end{equation}
where $\xi (0) =\xi$ and where for the different
GUTs the constant $b$ can be either positive,
negative or zero (see Ref. \cite{1}).
The gravitational running couplings are defined by the following
differential equations (we shall consider the gravitational
equations in
the Euclidean region)
\begin{eqnarray}
\frac{da_1 (t)}{dt} &=& \frac{1}{(4\pi)^2} \left[ \xi (t) -
\frac{1}{6} \right]^2 \frac{N_s}{2}, \nonumber \\
\frac{da_2 (t)}{dt} &=& \frac{1}{120(4\pi)^2} \left( N_s+6N_f+
12N_A \right), \nonumber \\
\frac{da_3 (t)}{dt} &=& - \frac{1}{360(4\pi)^2} \left( N_s+11N_f+
62N_A \right),
\label{544}
\end{eqnarray}
where $N_s,N_f$ and $N_A$ are the number of real
scalars, Dirac spinors and vectors, respectively (notice that the
running of $a_4(t)$ will not be meaningful for us, as we shall
see
below).
Owing to the fact that the theory under discussion is
multiplicatively
renormalizable, the effective Lagrangian satisfies the RG
equation
\begin{equation}
\left( \mu \frac{\partial}{\partial \mu} +
\beta_i \frac{\partial}{\partial \lambda_i} -
\gamma_i \phi_i \frac{\partial}{\partial \phi_i} \right)
L_{eff} (\mu,g_{\mu\nu}, \lambda_i, \phi_i )=0,
\label{545}
\end{equation}
where $\mu $ is the mass parameter, $\lambda_i =(g^2,h^2, f, \xi,
a_1,a_2,a_3,a_4)$ is the set of all coupling constants, the
$\beta_i$
are the corresponding $\beta$-functions and $\phi_i =(A_\mu,
\phi,
\psi)$.
The solution of Eqs. (\ref{545}) by the method of the
characteristics
gives (for all quantum fields we consider a zero background
field,
$\phi_i =0$):
\begin{equation}
L_{eff} (\mu,g_{\mu\nu}, \lambda_i)=
L_{eff} (\mu \, e^t,g_{\mu\nu}, \lambda_i (t)),
\label{546}
\end{equation}
where
\begin{equation}
\frac{d\lambda_i (t)}{dt} = \beta_i \left( \lambda_i (t)\right),
\ \ \ \
\ \ \lambda_i (0) =\lambda_i.
\label{547}
\end{equation}
Observe that for some of the coupling constants, the
corresponding
Eq. (\ref{547}) has been written above explicitly (Eq.
(\ref{544})),
while for a subset of them Eqs. (\ref{547}) have been actually
solved
(see Eqs. (\ref{542}) and (\ref{543})).
Actually, the idea itself of a
RG improvement procedure was suggested many years ago \cite{13}.
What we
do here is to make use once more of this interesting concept.
Physically, the meaning of expression (\ref{546}) is the
following:
$L_{eff}$ (called sometimes the Wilsonian effective action
\cite{5}) is
obtained through the above equations provided its functional form
for
some value of $t$ is known (usually it is the classical
Lagrangian that
serves as boundary condition at $t=0$). Another difficulty is
related
with the choice of $t$, which cannot be given a unique definition
due to
the presence, in general, of several different efective masses
(see the
discussions in Refs. \cite{6,7} concerning that point, for curved
and
for flat spaces, respectively).
There are different approaches to the gravitational effective
action
(for a general introduction, see \cite{8,1}). In the literature,
mainly
the case of a local effective action has been discussed (i.e.,
the
situation where the gravitational field is slowly varying).
One-loop
non-local effective actions have been considered in Refs.
\cite{9}-\cite{12} (see also the references therein), in
different
contexts, but almost exclusively the case of a free scalar field
theory
has been taken into account.
We will be interested in the situation where the gravitational
field is
weak, but rapidly varying, e.g.
\begin{equation}
\nabla \nabla R >> R^2.
\label{548}
\end{equation}
The non-local one-loop effective action for a free scalar field
theory
in this case has been calculated in Ref. \cite{10} (see also
\cite{11,12,14}), up to the second order on curvature
invariants.
Such a calculation is quite tedious, moreover, its extension to
other
fields (especially, to interacting fields) is anything but
trivial (see
\cite{11} for a discussion and list of references).
We will make use of this RG improvement technique in our
calculation,
what is going to yield a correspondingly more precise result than
the
one that has been obtained till now by means of previous
approaches to
the problem. First, all those calculations have been carried out
in the
one-loop approximation, while ours here will yield the RG improved
effective
Lagrangian to leading-log order (through summation of all possible
logarithms) of
perturbation theory, i.e., clearly beyond one-loop. Secondly, the
theory
under discussion had been usually restricted to scalar fields,
while the
considerations here will be applicable to any renormalizable
theory on a
curved background, including the ordinary renormalizable models
of
quantum gravity, as R$^2$-gravity (see \cite{1} for a review). In
particular we will present results for an arbitrary
asymptotically free
GUT in curved spacetime (see \cite{1,2}).
To begin, using the general expression (\ref{546}) we can
write
explicitly the RG improved effective Lagrangian for the theory
(\ref{541}), employing the classical Lagrangian as boundary
condition:
\begin{equation}
L_{eff} = a_1(t) R^2+ a_2(t) C^2_{\mu\nu\alpha\beta} + a_3(t) G +
a_4(t)
\Box R,
\label{549}
\end{equation}
where the choice of RG parameter $t$ will be described below.
{}From the explicit one-loop calculation \cite{4,2}, the RG parameter
is found to be
\begin{equation}
t \sim {1 \over 2}\ln {-\Box + c_1 R \over \mu^2 },
\label{5410}
\end{equation}
where the constant $c_1$ is different in the different sectors
(scalar,
spinor and vector). By looking at (\ref{548}) one can see that
in order to
get the dominant contribution we may just keep the first term in
(\ref{5410}), i.e. $t\simeq (1/2) \ln (-\Box /\mu^2)$. From the
explicit
study of the non-local effective action \cite{10}-\cite{12} it
follows
that
the thing one has to understand is the way non-local form factors
act,
as formal operators obeying the variational rules of finite
matrices (in
the Lorentzian region). Note also that the terms $a_4(t) \Box R$
and
$a_3(t) G$ are still total derivatives after the RG improvement
(compare with the other regime in \cite{6} where these terms
become
important). Notice that a different way of understanding the appearance
of the $-\Box$ under the logarithm is to resort to RG considerations in
curved space \cite{1}, where we know that a scale transformations of
the metric, $g_{\mu\nu} \rightarrow e^{-2t} g_{\mu\nu}$, ought to be
performed. Since, under this transformation, $R^2 \rightarrow e^{4t}
R^2$ and $\Box \rightarrow e^{2t} \Box$, the logarithm corresponding to
those terms becomes relevant in the high-energy limit $t\rightarrow
\infty$.
Finally, the RG improved non-local gravitational effective
Lagrangian
takes the form
\begin{eqnarray}
L_{eff} &=& R \left\{ a_1 - \frac{(\xi - 1/6)^2 N_s}{2B^2g^2
(2b+1)}
\left[
\left( 1 + \frac{B^2g^2 \ln (- \Box /\mu^2)}{2(4\pi)^2}
\right)^{2b+1}
-1 \right] \right\} R \nonumber \\ &&+
C_{\mu\nu\alpha\beta} \left[ a_2 + \frac{\ln (-\Box /\mu^2)}{240
(4\pi)^2} (N_s + 6N_f + 12N_A) \right]
C^{\mu\nu\alpha\beta},
\label{5411}
\end{eqnarray}
where $a_1$ and $a_2$ are intital values for the corresponding
effective
couplings. Notice that with the above form factors the solutions
fulfill
the requeriment of asymptotic flateness \cite{14}. As it has been
discussed in Refs. \cite{14,15}, the coefficients of the terms
linear in $\ln (-\Box)$ give a measure of the energy
radiation through the future null infinity.
Here we have obtained an effective Lagrangian, $L_{eff}$, which
sums
{\it all} the logarithms of perturbation theory, up to second
order
terms on curvature invariants on the background, of weak but
quickly
varying curvature. The theory under consideration is an
asymptotically
free GUT but, in principle, we can consider in the same way any
other
kind of renormalizable quantum field theory.
Notice, however, that the price one has to pay for the
universality of the approach (i.e., for the possibility to write
(\ref{5411}) for a variety of theories beyond the one-loop
approximation) is the fact that we cannot proceed to higher
orders in the curvature. The reason is that the terms as $R^3$,
$R^4, \ldots$ are ultraviolet finite. At the same time, the
ordinary technique to one-loop order \cite{11,12} gives the
possibility, in principle, to calculate the non-local effective
action up to any desired order in the curvature ---although it is
quite complicated, already in the case of the scalar theory.
It turns out, therefore, that the two approaches complement each
other quite well.
Using $L_{eff}$ one can obtain the effective gravitational
equations. Adding the quantum matter-induced effective Lagrangian
(\ref{5411}) to the classical Einstein Lagrangian (without the
cosmological constant, for simplicity), one gets the effective
gravitational equations in close analogy with Refs.
\cite{11,12,14}. Before doing this, it is convenient to rewrite
\begin{equation}
C_{\mu\nu\alpha\beta}^2 = G + 2 R_{\mu\nu}^2 - \frac{2}{3} R^2,
\label{5412}
\end{equation}
and to substitute it into Eq. (\ref{5411}). Then, one finds the
following Euclidean effective gravitational equations
\begin{eqnarray}
&&- \frac{1}{8\pi G} \left( R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} R
\right) + \left\{ a_1 + \frac{(\xi - 1/6)^2 N_s}{2B^2g^2 (2b+1)}
\left[
\left( 1 + \frac{B^2g^2 \ln (- \Box /\mu^2)}{2(4\pi)^2}
\right)^{2b+1}-1 \right] \right. \nonumber \\ &&-
\left. \frac{2}{3} a_2 - \frac{\ln (-\Box /\mu^2)}{360
(4\pi)^2} (N_s + 6N_f + 12N_A) \right\} \left[
4\nabla_\mu\nabla_\nu R - 4g_{\mu\nu} \Box R + {\cal O} (R^2)
\right] \label{5413} \\
&& + 2 \left[ a_2 + \frac{\ln (-\Box /\mu^2)}{240
(4\pi)^2} (N_s + 6N_f + 12N_A) \right]
\left[ 2\nabla_\mu\nabla_\nu R - g_{\mu\nu} \Box R -2 \Box
R_{\mu\nu}+ {\cal O} (R^2) \right] =0. \nonumber
\end{eqnarray}
Observe that in order to obtain the effective gravitational
equations it is not necessary to take into account the
$g_{\mu\nu}$-dependence of the form factors. As was discussed in
Ref. \cite{14}, the effective gravitational equations can be used
in order to study the problem of collapse.
To be remarked is the fact that the above approach works well for
renormalizable models of quantum
gravity too. In order to exemplify this, let us consider
R$^2$-gravity under the form
\begin{equation}
L = \frac{1}{\lambda} \left( R_{\mu\nu} - \frac{1}{3} R^2 \right) -
\frac{\omega}{3\lambda} R^2.
\label{5414}
\end{equation}
Such a theory is multiplicatively renormalizable, being non-unitary in
the perturbative approach (for a general review and a list of
references, see \cite{1}). The RG improved non-local effective
Lagrangian corresponding to this theory, with the same gravitational
background (\ref{548}), can be easily constructed. The effective
gravitational equations are (for simplicity, only leading-log terms
have been kept)
\begin{eqnarray}
&&\left[ \frac{1}{\lambda} +
\frac{133}{20 (4\pi)^2} \ln (- \Box /\mu^2)
\right]
\left[
\nabla_\mu\nabla_\nu R - g_{\mu\nu} \Box R - 2 \Box R_{\mu\nu} + {\cal
O} (R^2) \right] \nonumber \\
&& + 2 \left[
- \frac{1}{3\lambda}
- \frac{\omega}{3\lambda}
+ \left( \frac{10}{9} \omega^2 + \frac{5}{3} \omega + \frac{5}{36}
\right) \frac{\ln (-\Box /\mu^2)}{2(4\pi)^2}
\right] \nonumber \\ && \hspace{6mm} \times \left[ 4\nabla_\mu\nabla_\nu R
-4g_{\mu\nu} \Box R + {\cal O} (R^2) \right] =0,
\label{5415}
\end{eqnarray}
where $\lambda$ and $\omega$ are the initial values for the corresponding
effective couplings.
Following now Refs. \cite{14,15} (it is explained there in which form
non-local effective action can be relevant for black-hole physics), we
can discuss the implications that the above non-local gravitational
action has concerning the flux of the vacuum radiation in an
asymptotically free GUT. Working with the asymptotically flat
(Lorentzian) solution of Eqs. (\ref{5413}) one may consider the
congruence $u(x)=$ const. of the light rays that can reach the future
null infinity $\cal{F}^+$. We shall denote, as in \cite{14}, by $r$ the
luminosity distance along rays and by $M(u)$ the Bondi mass at
$\cal{F}^+$ (see \cite{17}). Then, the final expression for the
radiation corresponding to the vacuum energy in a spherically symmetric
state has been found to be the following \cite{14}:
\begin{equation}
\frac{d M(u)}{d u} = -\frac{1}{4\pi} (w_1+2w_2) \frac{d^2}{d^2u}
\int_{{\cal F}^-}^{{\cal F}^+} dr \, r \, R + {\cal O} (R^2),
\label{5416}
\end{equation}
where $w_1$ and $w_2$ are the coefficients of terms linear in $\ln
(-\Box)$ of (\ref{5411}), that is
\begin{equation}
L_{eff} = \left\{ R_{\mu\nu} \left[ a_1 - \frac{2}{3} a_2 -
\frac{w_1}{2(4\pi)^2} \ln \left( - \frac{\Box}{\mu^2} \right) \right]
R^{\mu\nu} + R \left[ 2a_2 -
\frac{w_2}{2(4\pi)^2} \ln \left( - \frac{\Box}{\mu^2} \right) \right] R
\right\},
\label{5417}
\end{equation}
where $a_1$ and $a_2$ can be taken to be zero and where $a_1(t)$ has
been expanded up to terms linear on $\ln (-\Box)$. Taking into account
the overall change of sign of $L_{eff}$ in the Lorentzian region, from
(\ref{544}) we obtain
\begin{equation}
w_1 = \frac{1}{60} \left( N_s + 6N_f + 12 N_A \right) , \ \ \ \
w_2 =-\frac{1}{180} \left( N_s + 6N_f + 12 N_A \right)+ \frac{N_s}{2}
\left( \xi - \frac{1}{6} \right).
\label{5418}
\end{equation}
In this way we can calculate the rate of the vacuum energy radiation
through the future null infinity, taking into account corrections to the
GUT under consideration.
To be remarked is the fact that the choice of $\xi$ can influence this
rate of radiation significantly (\ref{5418}). Radiation disappears when the
null surface $u=$ const. comes very close to the horizon \cite{14,15}.
Then, in order to find the Hawking radiation \cite{16} one has to
calculate the next-to-leading correction in (\ref{5417}), namely the
${\cal O} (R^2)$-terms.
To summarize, using rather simple RG considerations, we have constructed
a RG improved non-local gravitational Lagrangian corresponding to a
general asymptotically free GUT and also to R$^2$-quantum gravity. The
corresponding effective gravitational equations have been written
down as well. It would be now of interest to study the applications of
these equations to black hole physics in more detail, since they
certainly modify a number of results obtained previously.
\vspace{5mm}
\noindent{\large \bf Acknowledgments}
We would like to thank Roberto Percacci and Sergio Zerbini for their
interest in this work.
SDO is grateful with the members of the Department ECM, Barcelona
University, for warm hospitality.
This work has been supported by DGICYT (Spain), project
Nos. PB93-0035 and SAB93-0024, by CIRIT (Generalitat de
Catalunya), and by RFFR, project 94-020324.
\newpage
|
\section{\bf Introduction}
\setcounter{equation}{0}
The main goal of this paper is to study the topological space
of real {Lie} algebras of a given dimension $n\leq 4$.
Extensive studies have been dedicated to generalizations
of the classical {Lie} algebra structure. As an example think
of the famous q-deformations or Santilli's Lie isotopic
liftings \San. However,
few work has been dedicated to pursue the theory of
deformations and contractions of {Lie} algebras (or groups)
within their category.
{Smrz} \Smrz\ has considered the deformation of {Lie}
algebras outside a fixed subgroup. This kind of deformation
is in some sense complementary to a {In\"on\"u-Wigner} contraction
\In, which consists in a parametric linear and isotropic contraction
outside a given subgroup of a {Lie} algebra.
A particularly interesting problem is to find all possible
contractions and, more generally, all possible limit transitions
between real or complex {Lie} algebras of fixed
dimension $n$, and to uncover
the natural topological structure
of the space of all such {Lie} algebras.
It is clear that this requires, as a precondition,
to find all isomorphism classes of {Lie} algebras in the given
dimension.
Unfortunately, with increasing dimension $n$ the classification of
real and complex {Lie} algebras
becomes rapidly more complicated.
For this goal, the {Levi} decomposition into a semidirect
sum of a radical and a semisimple subalgebra proves to be useful.
This way {Turkowski} has classified
real {Lie} algebras which admit a nontrivial {Levi}
decomposition, up to $n=8$ in \aTur\
and recently for $n=9$ in \cTur.
In any dimension $n$, the classification of all
nilpotent {Lie} algebras is an essential step required for a complete
classification.
For $n=7$, a complete list of all nilpotent, real and complex,
{Lie} algebras has been given by {Romdhani} \Ro;
the complex case has been considered
first by {Ancochea-Bermudez} and {Goze} \An;
complex decomposable algebras have been studied by
{Charles} and {Diakite} \Ch.
The variety of structure constants of complex {Lie} algebras
has been examined for $n=4, 5, 6$ by {Kirillov} and {Neretin} \Ki.
{Grunewald} and {O'Halloran} \Gru\ have investigated the
complex, nilpotent {Lie} algebras for $n\leq 6$.
For $n=6$, all real nilpotent {Lie} algebras are classified
by {Morozov} \Mo;
solvable, non nilpotent
{Lie} algebras have been classified by {Mubarakzjanov} \cMu;
and solvable real {Lie} algebras containing nilradicals are
classified by {Turkovski} \bTur, thus completing
the classification of the solvable ones.
Both give reference
to the early
work of {Umlauf} \Um\
already classifying the nilpotent complex $6$-dimensional
{Lie} algebras.
{Mubarakzjanov} also classified
real {Lie} algebras up to $n=5$ in \bMu.
In \aMu\ he treats the case of real $n=4$, giving reference
to the early works of {Lie} \bLie\ for complex algebras with
$n\leq 4$ and, for the real $3$-dimensional case, to
{Bianchi} \bBi\ and later equivalent classifications of
{Lee} \Lee\ and {Vranceanu} \Vra.
The $3$-dimensional real {Lie} algebras,
are given by the so called {Bianchi} types,
classified independently
first by S. {Lie} \Lie\ and then by L. {Bianchi} \Bi.
The original classification of {Bianchi} revealed
9 inequivalent types of $3$-parameter {Lie} groups $G_3$,
numbered usually by the Roman numbers $\I,\ldots,\IX$.
The types of number \VI\ and \VII\
are actually
$1$-parameter sets
of {Lie} algebras,
\VIh\ resp. \mbox{${\rm VII}_h$} ,
with $h\geq 0$ all inequivalent.
We will refer to the inequivalent
$3$-dimensional real {Lie} algebras as the {Bianchi} types.
Our choice of the parameter $h$ is according
to {Landau-Lifschitz} \Lan,
which agrees for \mbox{${\rm VII}_h$} with {Behr}'s choice
in {\Est}.
When the isomorphism classes of {Lie} algebras for a given
real (or complex) dimension are known in a given dimension,
one can start to compare their algebraic structure systematically
and find their algebraic characteristics, i.e. the invariants.
So {Paterea} and {Winternitz} \PaW\
determined subalgebra structures for real {Lie} algebras with $n\leq 4$.
{Grigore} and {Popp} \Gri\ developed a general classification
of subalgebras of {Lie} algebras with solvable ideal,
and invariants of real {Lie} algebras have been calculated
for $n\leq 5$ by {Patera, Sharp, Winternitz} and {Zassenhaus}
{\PaSWZ}.
But the algebraic properties of {Lie} algebras are also related to
the topological structure of the space of all {Lie} algebras
in a given dimension.
On the space of all structure constants of real {Lie} algebras
in $n$ dimensions
{Segal} has introduced
(see page 255 in \Se) the subspace topology induced from $\R^{n^3}$.
The space $K^n$ of all isomorphism classes of real
$n$-dimensional {Lie} algebras
under general linear isomorphisms ${\GL}(n)$ of their generators
has a natural weakly separating
(i.e. $T_0$, not $T_1$) non-{Hausdorff} topology $\kappa^n$,
induced as the quotient topology from the {Segal} topology
by the equivalence relation given on the structure constants
via the action of $\GL(n)$.
This topology has been discovered
and described explicitly by {Schmidt} for $n\leq 3$ in \aSch\ and
more generally in \bSch.
As a real vector space, a Lie algebra admits also a natural orientation.
By the exponential map, for any {Lie} algebra there exists an
associated
{Lie} group which similarly admits the corresponding orientation
as a differentiable manifold.
Note that throughout the following any index or property concerning
orientation is set in brackets $()$ iff the corresponding quantity
can be considered optionally with or without reference to an orientation.
The present paper is organized by the following sections.
Sec. 2 resumes some well-known facts on {Lie} algebras
and topology needed in the sequel.
Sec. 3 describes the general construction of the topological spaces
$(K^n,{\kappa}^n)$ and $(K^n_{or},{\kappa}^n_{or})$, respectively
with and without orientation of the {Lie} algebras
as vector spaces.
Sec. 4 shows how those solvable elements
of $K^n$ which contain all the same ideal $J_{n-1}$ can be characterized
against each other by the normalized version (NJNF) of the
{Jordan} normal
form (JNF) of a single structure matrix.
Correspondingly, an oriented normalized {Jordan} normal form
(ONJNF) for the structure constants of oriented {Lie} algebras is defined.
Hence transitions $A\to B$ between {Lie} algebras can be described by
transitions between the corresponding normal forms.
Sec. 5 resumes important general properties
(see also {Schmidt} \bSch) of the
topology ${\kappa}^n_{(or)}$ and shows up further features of orientation
duality for arbitrary dimension $n$.
The structure of $K^n_{or}$, the space of equivalence classes of oriented
{Lie} algebras, as compared to its unoriented counterpart $K^n$,
has also been described in {Rainer} \aRa.
A generalization
of {Schmidt}'s notion of atoms
is made for arbitrary subsets of
$K^n_{(or)}$.
This is applied to the case of the non-selfdual subset
$K^n_{or}\setminus K^n_{SD}$,
decomposing it for $n=3$ and $n=4$
into its connected components $K^n_+$ and $K^n_-$.
Sec. 6 is devoted to the topology of the non oriented $K^n$ for $n\leq 4$.
The topological structure for $n\leq 4$ has also been described
by {Rainer} \aRa.
The $T_0$
topology $\kappa^n$ provides for $n\geq 3$ a rich local structure of $K^n$,
which we describe for $n\leq 4$.
In Sec. 6.1 the topological structure of $K^n$ for
$n\leq 3$ is analysed by use of the NJNF. So, using a quite different
method, we reproduce the results of
{Schmidt} in \aSch\ and \bSch.
Sec. 6.2 presents the detailed analysis of
the components of
$K^4$, their possible $\kappa^4$ limits, and transitions between them.
Thereby the relation between the
different classification schemes of
{Mubarakzjanov} \aMu, {Patera, Winternitz} \PaW\ and
{Petrov} \Pe\ is clarified.
Sec. 6.3
determines the topological structure of $K^4$.
Its parametrically connected components are related in a transitive
network of $\kappa^4$ transitions.
Sec. 7 is devoted to the topology of the oriented $K^n_{or}$ for $n\leq 4$,
which is also described
in {Rainer} \bRa.
In Sec. 7.1 the topological structure of $K^n_{or}$ for
$n\leq 3$ is analysed by use of the ONJNF, in correspondence with
results listed by {Schmidt} \bSch.
We give the connected components
$K^3_\pm$ explicitly.
Using the same method, Sec. 7.2 examines the orientation duality structure
of $K^4_{or}$ in detail. In particular,
we determine the connected components
$K^4_\pm$.
In Sec. 8 we discuss the present results.
\section{\bf Preliminaries}
\setcounter{equation}{0}
In the following we remind
shortly some of the notions needed throughout this paper.
A (finite-dimensional) {Lie} algebra is a
(finite-dimensional) vector space $V$, equipped with a
skew symmetric bilinear product
$[\cdot,\cdot]$ called {Lie} bracket,
which maps $(X,Y)\in V\times V$ to $[X,Y]=-[Y,X]\in V$ and satisfies
$\sum_{cycl.\atop X,Y,Z} [[X,Y],Z]=0, \forall X,Y,Z \in V$. The dimension of
the {Lie} algebra is the dimension of the underlying vector space.
Here and in the following all {Lie} algebras and vector spaces
are assumed to be finite-dimensional.
If the vector space is real resp.
complex, we say that the {Lie} algebra is real resp. complex.
If nothing else is specified in the following a {Lie} algebra or a
vector space is assumed to be real.
For a {Lie}
algebra $A$
the descending central series
of ideals is defined recursively by
\begin{equation}
C^0A:=A \quad{\rm and}\quad C^{i+1} A:=[A,C^{i}A]\subseteq C^{i}A.
\end{equation}
$A$ is called {\em nilpotent},
iff there exists a $p\in \N$, such that $C^pA=0$, i.e.
the descending central series
of ideals terminates at the zero ideal.
Furthermore for a {Lie} algebra $A$
the derivative series of ideals is defined
recursively by
\begin{equation}
A^{(0)}:=A \quad{\rm and}\quad A^{(i+1)}:=[A^{(i)},A^{(i)}]\subseteq A^{(i)}.
\end{equation}
$A$ is called {\em solvable},
iff there exists a
finite $q\in \N$, such that $A^{(q)}=0$, i.e. the derivative
series of ideals terminates at the zero ideal.
In the following, we consider
real {Lie} algebras of fixed finite dimension
$n\geq 2$ (for $n=1$ there is only 1 type of {Lie} algebra, namely
the {Abel}ian $A_1$), classified up to equivalence via real $\GL(n)$
transformations
of their linear generators $\{e_i\}_{i=1,\ldots,n}$,
which span an $n$-dimensional real vector space, which may in the following
be identified with $\R^n$ or the tangent space $T_xM$ at any point $x$ of an
$n$-dimensional smooth real manifold $M$. The {Lie} bracket $[\ ,\ ]$ is
given by its action on the generators $e_i$,
which is encoded in the structure constants $C^k_{ij}$,
\begin{equation}
[e_i,e_j]=C^k_{ij} e_k.
\end{equation}
(The sum convention is always understood implicitly, unless stated
otherwise.)
The bracket $[\ ,\ ]$ defines a {Lie} algebra, iff the structure constants
satisfy the $n\{{n\choose 2}+{n\choose 1}\}$ antisymmetry
conditions
\begin{equation}
C^k_{[ij]}=0,
\end{equation}
and the $n\cdot{n\choose 3}$ quadratic compatibility constraints
\begin{equation}
C^l_{[ij}C^m_{k]l}=0
\end{equation}
with nondegenerate antisymmetric indices $i,j,k$.
Here $_{[\quad ]}$ denotes
antisymmetrization w.r.t. the indices included.
Note that Eq. (2.5) is satisfied automatically by Eq. (2.4), if
the bracket is derived via $[e_i,e_j]\equiv e_i\cdot e_j-e_j\cdot e_i$
from an associative multiplication $e_i\cdot e_j$. In this case
Eq. (2.5) is an {\em identity}, called {Jacobi} identity.
Otherwise Eq. (2.5) is an {\em axiom}, which might be called
{Jacobi} axiom.
If there is an (adjoint) matrix representation of the algebra,
it is associative and hence
satisfies the {Jacobi} axiom (2.5) trivially, i.e. as identity.
We will not assume the existence
of any matrix representation nor any associative algebra multiplication,
because we want all the data for a {Lie} algebra to be encoded in the
structure constants. Hence we take (2.5) as an axiom.
The space of all sets $\{C^k_{ij}\}$ satisfying the {Lie} algebra
conditions (2.4) and (2.5)
can be viewed as a subvariety $W^n \subset \R^{n^3}$ of dimension
\begin{equation}
\dim W^n \leq n^3 - \frac{n^2(n+1)}{2} =\frac{n^2(n-1)}{2}.
\end{equation}
For $n=3$ the structure constants can be written as
\begin{equation}
C^k_{ij}= \mbox{$\varepsilon$} _{ijl}(n^{lk}+ \mbox{$\varepsilon$} ^{lkm}a_m),
\end{equation}
where $n^{ij}$ is symmetric and $ \mbox{$\varepsilon$} _{ijk}= \mbox{$\varepsilon$} ^{ijk}$ totally
antisymmetric with $ \mbox{$\varepsilon$} _{123}=1$.
With Eq. (2.7) the constraints
Eq. (2.5) are equivalent to
\begin{equation}
n^{lm}a_m=0,
\end{equation}
which are $3$ independent relations.
Actually, {Behr} has first classified the {Lie} algebras in $K^3$
according to their possible inequivalent eigenvalues of $n^{lm}$ and
values of $a_m$
(see {Landau-Lifschitz} \Lan).
With Eq. (2.8) also Eq. (2.5) is nontrivial for $n=3$.
Therefore the inequality in Eq. (2.6) is strict for $n\geq 3$.
Throughout the following, we will need the {\em separation axioms}
from topology (for further reference see also {Rinow} \Ri).
A given topology on a space $X$ is {\em separating} with increasing
strength if it satisfies one or more of the following axioms.
{\hfill \break}
{\bf Axiom $T_0$}: For each pair of different points there is an open set
containing only one of both.
\hfill\mbox\break
{\hfill \break}
{\bf Axiom $T_1$}: Each pair of different points has a pair of open
neighbourhoods with their intersection containing none of both points.
\hfill\mbox\break
{\hfill \break}
{\bf Axiom $T_2$} ({Hausdorff}):
Each pair of different points has a pair of disjoint neighbourhoods.
\hfill$\Box$\break
It holds: $T_2 {\Rightarrow } T_1 {\Rightarrow } T_0$. If a topology is only $T_0$, but not $T_1$,
we say that it is only {\em weakly separating} and speak also shortly
of the {\em weak} topology.
(The present notion {\em weak} should not be confused with another
one from functional analysis, which is not meant here.
{\em Separability} of the topological space is defined here by the
separation (german: Trennung) axioms $T_0, T_1$ or $T_2$.
This should not be confused with a further notion related to
the existence of a countable dense subset.)
The separation axioms can equivalently be characterized in terms
of sequences and their limits.
{\bf Lemma.} For a topological space $X$ the following equivalences hold:
{\hfill \break}
a) $X$ is $T_0$ $ {\Leftrightarrow } $ For each pair of points there is a sequence converging
only to one of them.
{\hfill \break}
b) $X$ is $T_1$ $ {\Leftrightarrow } $ Each constant sequence has at most one limit.
{\hfill \break}
c) $X$ is $T_2$ ({Hausdorff}) $ {\Leftrightarrow } $ Each {Moore-Smith}-sequence has
at most one limit.
{\hfill \break}
\hfill\mbox\break
(As a generalization of an ordinary sequence, a {Moore-Smith} sequence
is a sequence indexed by a (directed) partially ordered set.)
$T_1$ is equivalent to the requirement that each one-point set is closed.
We define for the following
the real {\em dimension} of a set as the largest number $k$
such that a subset homeomorphic to $\R^k$ exists.
\section{\bf Spaces $K^n$ and $K^n_{or}$ of {Lie} algebras}
\setcounter{equation}{0}
The space of structure constants $W^n$ can also be considered as a subvariety
of the fibrespace of
the tensor bundle $\wedge^2T^*M\otimes TM$
over any point of some smooth $\GL(n)$-manifold $M$.
If $M$ is oriented, the structure group of its tangent vector bundle
$TM$ is reduced from $\GL(n)$ to its normal subgroup
\begin{equation}
\GL^+(n)=\{A\in \GL(n):\det A > 0\}.
\end{equation}
Then $W^n$ gets an additional structure induced from $\wedge^2T^*M\otimes TM$
by the orientation of $M$. {\hfill \break}
$\GL(n)$ basis transformations induce
$\GL(n)$ tensor transformations between equivalent structure constants.
\begin{equation}
C^k_{ij} \sim (A^{-1})^k_h\ C^h_{fg}\ A^f_i\ A^g_j \ \ \forall A \in \GL(n),
\end{equation}
where $\sim$ denotes the equivalence relation.
This induces the space
\begin{equation}
K^n=W^n/\GL(n)
\end{equation}
of equivalence classes w.r.t. the
nonlinear action of $\GL(n)$ on $W^n$. The analogous space for the oriented
case is
\begin{equation}
K^n_{or}=W^n/\GL^+(n).
\end{equation}
The ${\GL}(n)$ action on $W^n$ is not free in general. It holds:
\begin{equation}
\dim W^n> \dim K^n_{(or)}\geq \dim W^n - n^2.
\end{equation}
The first inequality in Eq. (3.5) is a strict one,
because the (positive) multiples of the unit matrix in $\GL^{(+)}(n)$ give
rise to equivalent points of $K^n_{(or)}$.
Eqs. (2.6) and (3.5) provide only insufficient information on
$\dim K^n$. The latter is still unknown for general $n$. (For the
analogous complex varieties {Neretin} \Ne\ has given
an upper bound estimate.)
Let $\phi_{(or)}: W^n\to K^n_{(or)}$ be the canonical map for the
equivalence relation $\sim$ defined by the action of $\GL^{(+)}(n)$
in $W^n$.
The natural topology $\kappa^n_{(or)}$ of $K^n_{(or)}$ is given as the
quotient topology of the induced subspace topology of $W^n \subset \R^{n^3}$
w.r.t. the $\GL^{(+)}(n)$ equivalence relation.
In the oriented case, orientation reversal of the basis yields a
natural $Z_2$-action on $K^n_{or}$. This action is not free in general.
Hence the fibres of the projection
\begin{equation}
\pi: K^n_{or}\to K^n = W^n/{\GL}(n)=K^n_{or}/Z_2
\end{equation}
can be either $Z_2$ or $E$. In the first case
there is a pair of dual points, i.e. points that transform into each
other under the $Z_2$-action,
in the latter case it is a selfdual point in $K^n_{or}$.
The latter therefore
decomposes into a selfdual part $K^n_{SD}$, on which $Z_2$ acts
trivially, and 2 conjugate parts $K^n_{\pm}$. The latter are isomorphic to
each other by that reflection in $GL(n)$ that is chosen to define $Z_2$ in
Eq. (3.6).
\begin{equation}
K^n_{or}=K^n_{SD}\oplus K^n_{+} \oplus K^n_{-},
\end{equation}
where $\oplus$ denotes the disjoint union of subvarieties.
The projection $\pi$ has
the property that its restriction to $K^n_{SD}$ is the identity.
Therefore it is useful to make the following
{\hfill \break}
{\bf Definition 1.}
A point $A\in K^n$ is called {\em selfdual} if
$\pi^{-1}(A)\subset K^n_{or}$ consists of a single point, and
{\em non-selfdual}
if $\pi^{-1}(A)$ consists of a pair of dual points,
denoted by $A^R$ and $A^L$
respectively.
\hfill$\Box$\break
In order to yield a more explicit notion of selfduality, we formulate
{\hfill \break}
{\bf Lemma.}
{\em
A {Lie} algebra $A$ is selfdual, $A\in K^n_{SD}$,
if and only if there exist two different bases of $\R^n$ possessing different
orientation such that all the structure constants $C^k_{ij}$ concerning both
bases coincide.
}
\hfill$\Box$\break
Obviously a direct sum of a selfdual algebra with any other algebra
is selfdual.
Let us mention already here that
$K^n_{SD}$ is nonvoid for $n\geq 1$ while
$K^n_\pm$ are nonvoid sets only for $n\geq 3$.
We will see in Sec. 5 and 7 that the latter are actually nonvoid
for $n=3,4,5$ and at least any further odd $n$.
In any case $K^n_\pm$ are connected to $K^n_{SD}$.
We will see in Sec. 7 that each of $K^n_\pm$ is connected
for $n=3$ and $n=4$.
Note that for each pair of conjugate {Lie} algebras $A^R$ and $A^L$ it
is a priori completely arbitrary which one is assigned to $K^n_+$ and which
one to $K^n_-$. In order to reduce this arbitrariness, in Sec. 4 we will
minimize the number of connected components of $K^n_\pm$ to a single
component each, thus making $K^n_+$ and $K^n_-$ disconnected to each other.
However this requires first a better understanding of the
topological structure $K^n_{or}$.
When we do not want to care about effects of orientation,
instead of {Schmidt}'s topological space
$(G_n,\tau)\equiv (K^n_{or},\kappa^n_{or})$ from \bSch\
we will consider its projection to $(K^n,\kappa^n)$ by Eq. (3.6).
Let us define now the notion of transitions $A\to B$ in
$K^n_{(or)}$.
{\hfill \break}
{\bf Definition 2.}
Consider $A, B \in K^n_{(or)}$ with $A\neq B$.
If there is a sequence $\{A_i\}_{i\in \N}$ with
$A_i=A$ for all $i\in \N$ which for $i\to \infty$
converges to $B$ in the topology $\kappa^n_{(or)}$,
we say that there is a {\em transition} $A\to B$ in the topology
$\kappa^n_{(or)}$.
\hfill$\Box$ \break
Note that this definition makes sense because $K^n$ is a $T_0$
but not a $T_1$ space.
A transition is a special kind of limit characteristic for this topology.
{\hfill \break}
{\bf Convention.}
We distinguish in notation
between a concrete realization of a {Lie} algebra, $A$, and its
equivalence class, $[A]$, where ever this is relevant.
In the following, the former will an adjoint representation
of the latter, sometimes also called abstract,
{Lie} algebra.
However for notational simplicity we prefer to denote a point in $K^n$
by $A$ rather than by $[A]$. If the context does not give
the opportunity for confusion, $A$ is implicitly understood as
a shorthand for the (abstract) {Lie} algebra $[A]$.
\hfill$\Box$ \break
In the topology $\kappa^n_{(or)}$, a transition $A\to B$
occurs if and only if $B\in \cl \{A\}$.
For this transition the source $A$ is not closed, and the target
$B$ is not open in any subset of $K^n_{(or)}$ containing both of them.
In general, a point of $K^n_{(or)}$ will be neither open nor closed.
Open points only appear as a source, and not as a target, of transitions.
The structure of the rigid {Lie} algebras, which correspond just
to these open points, is examined in {Charles} \Cha.
Special kinds of transitions
on a certain 2-point set $\{A,B\}$
of {Lie} algebra isomorphism classes
are the contractions of {In\"on\"u-Wigner} \In\ and their
generalization by {Saletan} \Sal. For convenience let us define
these here.
Consider a $1$-parameter set of matrices $A_t\in\GL(n)$ with
$0<t\leq 1$, having a well defined matrix limit $A_0:=\lim_{t\to 0} A_t$
which is singular, i.e. $\det A_0=0$.
For given structure constants $C^k_{ij}$ of a {Lie} algebra
$A$ let us define for
$0<t\leq 1$ further structure constants
$C^k_{ij}(t):=(A^{-1}_t)^k_h\ C^h_{fg}\ (A_t)^f_i\ (A_t)^g_j$, which
according to (3.2) all describe the same {Lie} algebra $A$.
If there is a well defined limit $C^k_{ij}(0):=\lim_{t\to 0} C^k_{ij}(t)$
satisfying conditions (2.4) and (2.5)
then this limit defines structure constants of a {Lie} algebra $B$,
and the associated limit of {Lie} algebras $A\to B$
is called {\em contraction} according to {Saletan} \Sal\
or briefly {Saletan} {\em contraction}.
Note that a {Saletan} contraction $A\to B$ might yield either
$B=A$, then it is called {\em improper}, or $B\neq A$, then it is
a transition of {Lie} algebras.
A {Saletan} contraction is called
{In\"on\"u-Wigner} {\em contraction}
if there is a basis $\{e_i\}$ in which
$$
A(t)=
\left(
\begin{array}{cc}
E_m & 0 \\
0 & t\cdot E_{n-m}
\end{array}
\right)
\qquad \forall t\in [0,1],
$$
where $E_k$ denotes the $k$-dimensional unit matrix.
This definition closely follows {Conatser} \Co.
Given the latter decomposition, {In\"on\"u} and {Wigner} \In\ have
shown that the limit $C^k_{ij}(0)$ exists iff $e_i, i=1,\ldots,m$
span a subalgebra $W$ of $A$, which then characterizes the
contraction.
{Saletan} \Sal\ gives also a technical criterion for the
existence of the limit
$C^k_{ij}(0)$ defining his general contractions.
We only remark here that,
while a general {Saletan} contraction might be nontrivially iterated,
the iteration of an {In\"on\"u-Wigner} contraction is always improper,
i.e. no further contraction takes place.
Not every transition $A\to B$
corresponds to an {In\"on\"u-Wigner} contraction.
We will see some examples of transitions, which are given
only by a more general {Saletan} contraction \Sal.
However we will find also
transitions $A\to B$,
which are not even given by a {Saletan} contraction.
Transitions $A\to B$ in the topology $\kappa^n$ reveal for $n\geq 3$ a more
complicated structure of the underlying space $K^n$.
In $K^n$ transitions $A\to B$ and $B\to C$ imply a transition $A\to C$;
this means that transitions are transitive.
There is a partial order,
$A \geq B : {\Leftrightarrow } B\in \cl \{A\} {\Leftrightarrow } A\to B$
(which is also called the {\em specialization order}), which gives
$K^n_{(or)}$ the structure of a transitive network of transitions.
Since {Saletan} contractions \Sal\ are not transitive they do not
exhaust all kinds of possible $\kappa^n$ transitions.
Given the topology of $K^n$, on any 2-point subset $\{X,Y\}\subset K^n$
we can take the induced topology and consider the set
$T^n:=\{\{X,Y\}\subset K^n\vert X\neq Y\}$ of all 2-point topological
subspaces of $K^n$.
Note that a $T_0$ topological space, like that of $K^n$ for $n\geq 3$,
is in general not determined by the set $T^n$ of all its induced
2-point topological subspaces.
However if the topological space under consideration
is finite then $T^n$ determines already its topology,
which is trivially true
for $K^1$ and $K^2$.
\section{\bf Normal forms of the structure constants}
\setcounter{equation}{0}
The structure constants of $A_n\in [A_n]\in K^n$ are given by
the $n$ matrices $C_i:=(C^k_{ij})$, $i=1,\ldots,n$,
with rows $k=1,\ldots,n$ and
columns $j=1,\ldots,n$. $C_i$ is just the matrix of ad$e_i$ w.r.t.
the basis $e_1,\ldots,e_n$. {\hfill \break}
By Eq. (2.4), the column $j=i$ vanishes identically $\forall C_{i}$.
Furthermore
the diagonals $(C^j_{ij})$, $i=1,\ldots,n$ (no j-summation),
determine the rows
with $k=i$, since $(C^i_{ij})=(-C^i_{ji})$, $i=1,\ldots,n$
(no i-summation).
Therefore $A_n$ is described completely by the
$(n-1)\times (n-1)$-matrices
$C_{<i>}:=(C^k_{ij})$, $i=1,\ldots,n$, with $k,j\neq i$ and
$1\leq k,j\leq n$. {\hfill \break}
In the special case where $A_n$ has an ideal
$J_{n-1}\in [J_{n-1}]\in K^{n-1}$,
we take without restriction $[A_n]/[J_{n-1}]=\mbox{span}(e_n)$.
Then $A_n$ with a given
$J_{n-1}$ is described completely by $C_{n}$ or $C_{<n>}$ only.
{\hfill \break}
{\bf Definition 3.} The {\em normalized} JNF (NJNF) of a matrix $C$ is given
by the {Jordan} normal form, abbreviated JNF,
of $C$ modulo $\R\setminus \{0\}$, i.e. given by the equivalence
class of JNFs, which differ only by a common absolute scale and a common
overall sign of their nonzero eigenvalues w.r.t. the eigenvalues of $C$.
(The {Jordan} block structure and
the multiplicities are the same for all of
them.)
\hfill $\Box$\break
Thus a normalization convention for the JNF is the division of all
nonzero eigenvalues by a fixed element of $\R {\setminus } \{0\}$. If not stated
otherwise,
we divide in the following just by the (absolutely) largest eigenvalue in order
to
represent the NJNF class of the JNF. {\hfill \break}
Note that the $n^{th}$ row and column of $C_n$
add only an additional
eigenvalue 0 (as {Jordan} block) to the JNF or NJNF of $C_{<n>}$.
Since absolute scaling of all eigenvalues of a structure matrix $C_{<n>}$
by $\lambda \in \R {\setminus } \{0\}$ can be
achieved by stretching the basis $\{e_i\}$ homogeneously by $\lambda^{-1}$,
it is an equivalence transformation of the algebra. On the other hand it is
evident that changing in $C_{<n>}$ the ratio $r$ of any 2 eigenvalues to
$r'$, such that $r'$ is not a ratio of any original
eigenvalues, changes the equivalence class.
{\bf Theorem.}
{\em
Consider the set of algebras $A_n$ which have a
common (abstract) ideal $J_{n-1}$.
Then $A^{(1)}_n\sim A^{(2)}_n$, iff the matrices $C^{(1)}_{<n>}$ and
$C^{(2)}_{<n>}$ have the same \NJNF.
} {\hfill \break}
{Proof:} $A^{(1)}_n\sim A^{(2)}_n$ iff $\exists M\in {\GL}(n):
{C^{(1)}}^k_{ij} = (M^{-1})^k_h\ {C^{(2)}}^h_{fg}\ M^f_i\ M^g_j
\sim M^f_i {C^{(2)}}^h_{fg}$. By linearity of $[\ ,\ ]$ in the second
argument, the linearly independent recombinations
${\tilde{C}}^{(2)}_i:=M^f_i\ C^{(2)}_{f}$ describe still the same algebra
as $C^{(2)}_{i}$. In particular, the (abstract) ideal $J_{n-1}$
is invariant under $M$.
Since the algebras have the same ideal $J_{n-1}$,
they are characterized by
the matrices $C^{(1)}_{<n>}$ resp. $C^{(2)}_{<n>}$. They describe
inequivalent algebras, iff $C^{(1)}_{<n>}$ is inequivalent (modulo overall
scaling by $M=\lambda E_n,\ \lambda \in \R {\setminus } \{0\}$) to
${\tilde{C}}^{(2)}_{<n>}$ and therefore also to $C^{(2)}_{<n>}$.
But the equivalence class of any
structure matrix $C_{<n>}$ is described by its (real) JNF modulo homogeneous
scaling of the eigenvalues with $\lambda \in \R {\setminus } \{0\}$.
\hfill $\Box$ \break
{\hfill \break}
Already {Mubarakzyanov} \aMu\ had realized the advantage given by
an ideal $J_{n-1}$ of codimension $1$. Since then also others, like
{Magnin} \Mag\ within the nilpotent {Lie} algebras of dimension $\leq 7$,
systematically cosidered subclasses of algebras which have a fixed
{Lie} algebra of codimension $1$.
In the following, we consider without restriction
of generality the ideals $J_{(n-1)}$
in the normal form given by the NJNF of the structure constants.
The equivalence class $[A_{n}]$ of any algebra $A_{n}$ with a
fixed normal class ideal and additional structure constants from $C_{<n>}$
will be characterized in the following by the NJNF of $C_{<n>}$
and denoted by
$$
\NJNF(A_n) := \NJNF(C_{<n>}).
$$
Now we can define the ONJNF of structure matrices $C_{<n>}$ of oriented
{Lie} algebras.
{\hfill \break}
{\bf Definition 4.}
The ONJNF of the structure matrix $C_{<n>}$ of an oriented {Lie} algebra
$A_n$ is set identical to its NJNF if $A_n$ is selfdual, and
it is given as
$ \mbox{${\rm ONJNF}$} (C_{<n>}):=\pm \NJNF(C_{<n>})$ for $A_n\in K^n_{\pm}$ respectively.
\hfill $\Box$\break
If $A_n$ is characterized by an ideal $J_{n-1}$ in normal form
then we set
$$
\mbox{${\rm ONJNF}$} (A_n) := \mbox{${\rm ONJNF}$} (C_{<n>}).
$$
\section{\bf General properties of $\kappa^n$ and $\kappa^n_{or}$}
\setcounter{equation}{0}
In this section we describe the general topological properties of the
topological space $(K^n_{(or)},\kappa^n_{(or)})$.
Let us first remind some general properties
from {Schmidt} \bSch\ (where also more details and proofs
can be found).
{\hfill \break}
{\bf Proposition.}
{\em
$K^n_{(or)}$ has the following properties w.r.t. $\kappa^n_{(or)}$:
}
{\hfill \break}
a)
{\em
The {Abel}ian algebra $\{nA_1\} \subset K^n_{(or)}$ is the only closed
1-point set and is contained in any nonempty closed subset of $K^n_{(or)}$.
}
{\hfill \break}
b)
{\em
$K^n_{(or)}$ is connected and compact.
}
{\hfill \break}
c)
{\em
$K^n_{*(,or)}:=K^n_{(or)} {\setminus } \{nA_1\}$ is a compact space, but
$K^n_{*(,or)}$ is not a
closed subset of $K^n_{(or)}$.
}
{\hfill \break}
d)
{\em
$K^n_{*(,or)}$ is {Hausdorff} $(T_2)$ for $n=2$ only.
}
{\hfill \break}
e)
{\em
For $n\geq 2$ (resp. $n\geq 3$) the separability of $K^n_{(or)}$
(resp. $K^n_{*(,or)}$) is only weak ($T_0$, i.e. for each pair of points
there is a sequence converging to only one of them).
}
\hfill$\Box$\break
d) and e) correspond to the fact that, though $K^n$ is still an algebraic
variety (defined by purely algebraic relations (2.4), (2.5) and
(3.2)), it can not be expected to be a (topological $T_1$)
manifold.
$K^n$ is the orbit space of $W^n$ w.r.t. the action of the
noncompact group ${\GL}(n)$, which behaves algebraically badly on $W^n$
for $n\geq 2$. So some of the orbits (the elements
of $K^n$) are closed in $K^n$, others are not.
Strong separability ($T_1$, i.e. each constant sequence has at most one
limit) would imply that there should not exist
transitions $A \to B$ between
inequivalent {Lie} algebra classes $A\not\sim B$, given by a sequence
$\{A_i\}$ of {Lie} algebras of class $A$ converging to a
{Lie} algebra of class $B$.
But this is exactly what happens for dimension $n\geq 2$, as will be
seen explicitly below. Obviously transitions $A\to B$ will be transitive,
which decisively effects the topology of $K^n$.
{\hfill \break}
Transitions which are impossible in a given dimension $n$
can become possible
after {Abel}ian embedding into dimension $n+1$. Therefore the following
lemma holds. {\hfill \break}
{\bf Lemma 1.}
{\em
The {Abel}ian embedding $\oplus \R$ of $K_n$
into $K_{n+1}$ is continuous, but for $n\geq 2$ not homeomorphic.
}
\hfill$\Box$ \break
So we are led to the following
{\hfill \break}
{\bf Definition 5.}
The {\em essential dimension} of an $n$-dimensional (oriented)
{Lie} algebra $A_n$ is
defined as the smallest possible number $n_e\leq n$, such that
$A_n=A_{n_e}\oplus \R^{n-n_e}$. The essential-dimensional
subset of $K^n$ is defined as $K^n_{de}=\{A\in K^n\vert n_e(A)=n\}$
\hfill$\Box$ \break
{\bf Lemma 2.}
{\em
The subsets $\{A\in K^n\vert n_e(A)\leq m\}$ for any
fixed $m\leq n$ need not to be closed.
}
\hfill$\Box$ \break
This is due to the existence of transitions or limits
of structure constants in NJNF
such that one or more NJNF eigenvalues degenerate to another one
(in Lemma 2 it is the eigenvalue $0$), initially distinct from them;
in this case the algebraic multiplicity of this eigenvalue increases
automatically, but its geometric multiplicity (expressed by its number of
{Jordan} blocks) may remain constant, since an eigenvector of
an eigenvalue different from the limit eigenvalue may converge to a
principal (not necessarily eigen) vector of the limit eigenvalue
($0$ for Lemma 2). {\hfill \break}
A {Lie} algebra characterized by structure constants $C^k_{ij}$ is
called {\em unimodular}
(on a corresponding {Lie} group) iff $ \mbox{${\rm tr }$} (C_i) = C^k_{ik} =0\
\forall i$, where the adjoint representation is generated by the matrices
$C_i$. We denote the subset of all points in $K^n$ that correspond
to unimodular {Lie} algebras by $U^n$, and set $U^n_*=U^n\cap K^n_*$.
Since the zero set of a continuous function is always closed, we have {\hfill \break}
{\bf Lemma 3.}
{\em
The unimodular subset $U^n\subset K^n$ is closed and compact.
}
\hfill$\Box$ \break
For $n\geq 2$ the structure constants of any {Lie} algebra admit,
as a tensor $C$, the irreducible decomposition \bSch
\begin{equation}
C^k_{ij}=D^k_{ij}+\delta^k_{[i}v_{j]}
\end{equation}
in a tracefree part $D$ with tensor components $D^k_{ij}$ (the trace free
condition for $D$ can be written as
$ \mbox{${\rm tr }$} (D_i)=D^k_{ik}=0\ \forall i$),
and a vector part, constructed from a vector $v$ with components
$v_i=C^j_{ij}/(1-n)$ and the Kronecker
symbol of components $\delta^k_i$ (remind the sum convention over upper and
lower indices and the convention to perform an antisymmetric sum over all
permutations of the indices included in $_{[\quad ]}$).
The {Lie} algebra is {\em unimodular}
(like any associated connected {Lie} group),
iff it is tracefree, $v\equiv 0$, and it is said to be of
{\em pure vector type},
iff $D\equiv 0$. In this sense the unimodular and pure vector type
are complementary.
The class ${\V}^{(n)}$ of pure vector type is selfdual for all $n\geq 2$.
It is the generalization of the unique non-{Abel}ian $2$-dimensional
algebra $A_2$ (see Sec. 6) to arbitrary $n$.
So for each $n$, there exists exactly
one non-{Abel}ian pure vector type {Lie} algebra, denoted by $\V^{(n)}$ because
for $n=3$ it is the Bianchi type \V.
It has the {Abel}ian ideal $\I^{(n-1)}$ and
$[\NJNF(\V^{(n)})]^k_j=\delta^k_j$.
For convenience we mention explicitly the nonvanishing commutators
of $\V^{(n)}$, for an adapted basis $\{e_1,\ldots,e_n\}$:
\begin{equation}
[e_n, e_i] =e_i,\ i=1,\ldots,{n-1}.
\end{equation}
The $3$-dimensional Heisenberg algebra (= Bianchi type \II)
is defined in its NJNF by the
nonvanishing commutators
\begin{equation}
[e_3, e_2] =e_1.
\end{equation}
By {Abel}ian embedding we define the
class ${\II}^{(n)}:={\II}\oplus \R^{n-3}$ for $n\geq 3$.
Like\II, it is unimodular and nilpotent of degree 2.
${\II}^{(n)}$ is non-selfdual for $n=3$ and selfdual for $n\geq 4$.
Its nonvanishing structure constants for an adapted basis
$\{e_1,\ldots,e_n\}$)
are given by Eq. (5.3)
with all indices increased by $n-3$.
In {Schmidt} \bSch, an element $A_n\in K^n_*$ for which its closure
in $K^n$ consists of 2 elements only, $\cl\{A_n\}=\{A_n,nA_1\}$,
was called an {\em atom}.
Here we will prefer to call equivalently $A_n$ an atom of $K^n_*$, iff
its closure in $K^n_*$ is $\cl_{K^n_*}\{A_n\}=\{A_n\}$.
Let us generalize this:
{\hfill \break}
{\bf Definition 6.}
For any subset $S\subset K^n_{(or)}$,
an element $A\in S$ is called an {\em $S$-atom},
iff it is closed w.r.t. $S$, i.e.
$\cl_S\{A\}=\{A\}$.
\hfill$\Box$ \break
In the following, we call an $S$-atom also synonymously
an {\em atom of $S$} and assume $S=K^n_*$ if not specified otherwise.
Recall
from {Schmidt} \bSch
{\hfill \break}
{\bf Theorem 1.}
{\em
For $n=2$ there is only 1 atom, $A_2\equiv {\V}^{(2)}$. {\hfill \break}
For each $n\geq 3$ there exist exactly 2 atoms, the unimodular ${\II}^{(n)}$
and the pure vector type ${\V}^{(n)}$.
}
\hfill$\Box$ \break
For $n\neq 3$ all atoms are selfdual.
If we consider the corresponding atoms of $K^n_{*,or}$, then only for
$n=3$ there is a difference to the nonoriented case. Instead of the
unique non-selfdual atom $\II$ in $K^3$, there exist 2 non-selfdual
atoms, $\II^R$ and $\II^L$, in $K^3_{or}$.
For each $n\geq 3$ there is an algebra ${\IV}^{(n)}$,
given by
$[\NJNF(\IV^{(n)})]^k_j=\delta^k_j+\delta^k_{n-2}\delta^{n-1}_j$ w.r.t. to
the {Abel}ian ideal $\I^{(n-1)}$. It is selfdual for $n\geq 4$ and
non-selfdual for $n=3$.
For convenience we mention explicitly the nonvanishing commutators
of $\IV^{(n)}$, for an adapted basis $\{e_1,\ldots,e_n\}$ given by
\begin{equation}
[e_n, e_i] =e_i,\ i=1,\ldots,n-2, \quad [e_n, e_{n-1}] =e_{n-2}+e_{n-1}.
\end{equation}
$K^n_*$ is generated by infinitesimal deformations of the atoms;
this means: $K^n_*$ itself is the only open subset of $K^n_*$
which contains all atoms.
Since both, ${\IV}^{(n)} \to {\II}^{(n)}$ and
${\IV}^{(n)} \to {\V}^{(n)}$, it follows that $K^n_*$ is connected. {\hfill \break}
Remark: Connectedness is trivial for $K^n$, but non-trivial for $K^n_*$.
To understand better where the exceptionality of $n=3$ w.r.t. to duality
comes from,
realize that $n_e(\V^{(n)})=n$ but $n_e(\II^{(n)})=3$
for all $n\geq 3$. In particular, $\II^{(n)}$ has essential dimension
$n_e=n$ only for $n=3$; for $n\geq 4$ it is decomposable and hence
selfdual.
More generally there holds
{\hfill \break}
{\bf Lemma 4.}
{\em
$K^n_{NSD}:=K^n {\setminus } K^n_{SD}$ is contained in the subset $K^n_{de}$
of $K^n$ for which $n_e=n$.
}
\hfill$\Box$ \break
To overcome the difference in the essential dimension of the atoms for
$n\geq 4$, let us search for atoms w.r.t. the subset $K^n_{de}$
of essential dimension $n_e=n$ in $K^n$. We find
{\hfill \break}
{\bf Theorem 3.}
{\em
The set $K^n_{de}$ has the following atoms:
{\hfill \break}
a)
For $n\geq 2$
exactly $1$ pure vector type atom, called $\ve(n)$.
}
{\hfill \break}
b)
{\em
For $n\geq 3$ a nilpotent unimodular atom, called $\ii(n)$,
located in the subset of algebras with ideal $\I^{(n-1)}$.
}
{\hfill \break}
c)
{\em
For $n\geq 5$ further $]\frac{2}{3}(n-4)[$ mixed type atoms,
denoted $a_m(n)$, $m=2+[\frac{n-4}{3}],\ldots, n-3$,
all located in the subset of algebras with ideal $\I^{(n-1)}$.
{\hfill \break}
(Here $[x]$ resp. $]x[$ denotes the largest/smallest integer
less/greater or equal than x.)
Within the subspace $K^n_{de\vert\I^{(n-1)}} \subset K^n_{de}$
given by $K^n_{de}$-algebras with ideal $\I^{(n-1)}$
there are no further $K^n_{de}$-atoms than that of {\rm a), b)} and
{\rm c)}.
$K^n_{de\vert\I^{(n-1)}}$ is connected.
}
{\hfill \break}
Proof:
a) By Theorem 1 the algebra $\V^{(n)}$ is an atom of $K^n_*$.
Since $K^n_{de}\subset K^n_*$ and $n_e(\V^{(n)})=n$,
it follows
that $\V^{(n)}$ is an atom of $K^n_{de}$. Any algebra with
only nonzero components $v_i$ in the vector $v$
of the decomposition (5.1) has a transition
or limit to $\V^{(n)}$.
Hence $\ve(n):=\V^{(n)}$ is the unique (pure) vector type $K^n_{de}$-atom.
b)
Some of the algebras with some vanishing component $v_i$
have transitions or limits to an algebra with $v\equiv 0$.
Hence we have to search for unimodular $K^n_{de}$-atoms of essential
dimension $n_e=n$. Such an atom is the nilpotent algebra
$\ii(n)$ with $\NJNF(\ii(n))$
w.r.t. the ideal $I^{(n-1)}$ given for even $n$
as a direct sum
of $1$ block of $\NJNF(\ii(4))$ and further blocks of $\NJNF(\ii(3))$,
and for odd $n$
as a direct sum of $\NJNF(\ii(3))$ blocks only,
where
$$
\NJNF(\ii(3)):=
\left(
\begin{array}{cc}
0 & 1 \\
& 0
\end{array}
\right)
$$
and
\begin{equation}
\NJNF(\ii(4)):=
\left(
\begin{array}{ccc}
0 & 1 & \\
& 0 & 1 \\
& & 0
\end{array}
\right).
\end{equation}
The algebra $\ii(n)$ is essential-dimensional,
because any of its subalgebras
$\ii(3)$ and $\ii(4)$ is so; it is an
$K^n_{de}$-atom, because its only possible limits
necessarily
generate a $1\times 1$-block $(0)$ in its NJNF, thus decreasing $n_e$
at least by $1$.
c)
The mixed atoms can be characterized by their NJNF w.r.t. the ideal
$\I^{(n-1)}$. Let us set
$$
\NJNF(a_m(n)):=\NJNF(\ve(m+1))\oplus\NJNF(\ii(n-m)).
$$
Since for
$m=2+[\frac{n-4}{3}],\ldots, n-3$ the geometric multiplicity $m$
(= the number of {Jordan} blocks) of the eigenvalue $1$ is
bigger than that of the eigenvalue $0$,
any transition yields an additional {Jordan} block
$0$ and hence leaves $K^n_{de}$. So, being essential-dimensional,
$a_m(n)$ is an $K^n_{de}$-atom for $m=2+[\frac{n-4}{3}],\ldots, n-3$.
Any algebra of $K^n_{de\vert\I^{(n-1)}}$ has a combination of
transitions and parametric limits leading
to at least one of the atoms from a), b) or c), depending on the
degeneracy of its eigenvalues. The only nontrivial case,
which remains to be checked, are the algebras with
their NJNF w.r.t. an ideal $\I^{(n-1)}$ given as
$\NJNF(\ve(m+1))\oplus\NJNF(\ii(n-m))$ where
$m=1, \ldots, 1+[\frac{n-4}{3}]$ and $n\geq 4$.
But any of these has a transition to $\ii(n)$.
Let us now consider some algebra in $K^n_{de\vert\I^{(n-1)}}$ with
only nondegenerate nonzero eigenvalues. By continuous deformation
of its eigenvalues, such that every deformed algebra remains in
$K^n_{de\vert\I^{(n-1)}}$, and transitions within
$K^n_{de\vert\I^{(n-1)}}$ each of the atoms a), b) and c)
can be reached. Since these have just been seen to be the only
atoms of $K^n_{de\vert\I^{(n-1)}}$ it follows that
$K^n_{de\vert\I^{(n-1)}}$ is connected.
\hfill$\Box$ \break
$K^n_{de}$ itself might have further atoms
located in $K^n_{de}\setminus K^n_{de\vert\I^{(n-1)}}$.
Since these are difficult to find, in general
one cannot see whether $K^n_{de}$ is connected.
The nonvanishing commutators
of $\ii(n)$, $n\geq 3$,
are given for an adapted basis $\{e_1,\ldots,e_n\}$
explicitly by
$$
[e_n, e_2] =e_{1},\ [e_n, e_3] =e_{2},\ i=2,\ldots,n-1.
$$
\begin{equation}
[e_n, e_{2i+3}] =e_{2i+2},\quad i=1,\ldots,\frac{n-4}{2}.
\end{equation}
for $n$ even and by
\begin{equation}
[e_n, e_{2i}] =e_{2i-1},\quad i=1,\ldots,\frac{n-1}{2}.
\end{equation}
for $n$ odd.
$\ii(3)\equiv \II$ is the {Heisenberg} algebra.
The number of $\NJNF(\ii(3))$ blocks in its NJNF is even for
$n\equiv 0 \,\mbox{mod}\, 4$ or $n\equiv 1 \,\mbox{mod}\, 4$, and it is
odd for $n\equiv 2 \,\mbox{mod}\, 4$ or $n\equiv 3 \,\mbox{mod}\, 4$.
The mixed types $a_m(n)$, $n\geq 5$,
have respective algebraic and geometric multiplicities
$m=2+[\frac{n-4}{3}],\ldots, n-3$
for the eigenvalue $1$.
Their nonvanishing commutators
are given
w.r.t. an adapted basis
{\hfill \break}
$\{e_1,\ldots,e_n\}$ as
$$
[e_n, e_1] =e_{1},\ldots, [e_n, e_m] =e_{m},
$$
$$
[e_n, e_{m+2}] =e_{m+1},\ [e_n, e_{m+3}] =e_{m+2},\quad i=2,\ldots,n-1,
$$
\begin{equation}
[e_n, e_{2i+m+3}] =e_{2i+m+2},\quad i=1,\ldots,\frac{n-m-4}{2},
\end{equation}
for $n-m$ even, and by
$$
[e_n, e_1] =e_{1},\ldots, [e_n, e_m] =e_{m},
$$
\begin{equation}
[e_n, e_{2i+m}] =e_{2i+m-1},\quad i=1,\ldots,\frac{n-m-1}{2},
\end{equation}
for $n-m$ odd.
The reflection $e_1\to -e_1$ leaves $\ve(n)$ and any mixed type
atom $a_m(n)$ invariant; hence all these atoms are selfdual.
The nilpotent
atom $\ii(n)$
remains as the only possibility for a non-selfdual $K^n_{de}$-atom
within $K^n_{de\vert\I^{(n-1)}}$.
Therefore, next we want to examine the orientation duality of $\ii(n)$.
{\hfill \break}
{\bf Theorem 4.}
{\em
For $n\geq 3$ the $K^n_{de}$-atom $\ii(n)$ is non-selfdual
only if $n\equiv 3\, \mbox{\rm mod}\, 4$.
{\hfill \break}
$\ii(n)$ non-selfdual for $n\equiv 3\, \mbox{\rm mod}\, 4$
implies that $\ii(n)$ is a $K^n_{de}$-atom.
} {\hfill \break}
{Proof:} A combination of the reflections $e_n\to-e_n$
and $e_{2i}\to-e_{2i}$ for $i=1,\ldots,[\frac{n-1}{2}]$
leaves $\ii(n)$ invariant.
The total number of these reflections is $[\frac{n+1}{2}]$,
which is odd for $n\equiv 1 \,\mbox{mod}\, 4$
or $n\equiv 2 \,\mbox{mod}\, 4$.
Furthermore for $n$ even, $e_i\to-e_i$, $i=1,\ldots,n-1$ yields
a reflection keeping $\ii(n)$ invariant.
So for all $n$ but $n\equiv 3 \,\mbox{mod}\, 4$ the algebra is selfdual.
Any limit of $\ii(n)$ is selfdual, because it is a $K^n_{de}$-atom
and any non-essential-dimensional algebra is decomposable and hence
selfdual.
Therefore non-selfduality for $n\equiv 3\, \mbox{\rm mod}\, 4$
implies that $\ii(n)$ is a $K^n_{de}$-atom.
\hfill$\Box$ \break
For $n\equiv 3 \,\mbox{mod}\, 4$ it was impossible to construct a reflection
leaving $\ii(n)$ invariant. But when there is no such reflection
the algebra is non-selfdual.
Let us define for $n\geq 3$ an algebra $\iv(n)$
given for an adapted basis $\{e_1,\ldots,e_n\}$ by
the nonvanishing commutators
$$
[e_n, e_1] =e_{1},\
[e_n, e_2] =e_{1}+e_{2},\ [e_n, e_3] =e_{2}+e_{3},\ i=2,\ldots,n-1,
$$
\begin{equation}
[e_n, e_{2i+3}] =e_{2i+2}+e_{2i+3},\ i=1,\ldots,\frac{n-4}{2},
\end{equation}
for $n$ even, and by
\begin{equation}
[e_n, e_{2i}] =e_{2i-1}+e_{2i},\ i=1,\ldots,\frac{n-1}{2},
\end{equation}
for $n$ odd.
By similar considerations as for $\ii(n)$ in Theorem $3$ one finds
that $\iv(n)$ is non-selfdual for only for $n$ odd and selfdual for
$n$ even.
In any case it has an ideal $\I^{(n-1)}$
and for $n$ odd the geometric multiplicity of its eigenvalue
$1$ of the NJNF w.r.t. $I^{(n-1)}$ can only be increased by yielding
at least two $1\times 1$ blocks of that eigenvalue, hence the
resulting algebra of such a transition is selfdual.
Apart from limits which increase multiplicity, the only further limits
of $\iv(n)$ are transitions with the eigenvalue becoming $0$,
either to $\ii(n)$ or some limit thereof.
But, according to
Theorem 4, for $n\not\equiv\,\mbox{mod}\,4$, the algebra $\ii(n)$ is selfdual.
Any limits of $\ii(n)$ are selfdual, because it is a $K^n_{de}$-atom
and any non-essential-dimensional algebra is decomposable and hence
selfdual.
Hence non-selfduality of $\iv(n)$ for $n$ odd implies
that $\iv(n)$ is a $K^n_{NSD}$-atom for $n\equiv 1 \,\mbox{mod}\, 4$.
Non-selfduality of $\ii(n)$ for $n\equiv 3\, \mbox{\rm mod}\, 4$
implies further that $\iv(n)$ is
no atom for $n\equiv 3\, \mbox{\rm mod}\, 4$.
If for $n\equiv 3\, \mbox{\rm mod}\, 4$ resp. $n$ odd the algebras
$\ii(n)$ resp. $\iv(n)$ are in fact non-selfdual,
we get the
{\hfill \break}
{\bf Corollary.}
{\em
For odd $n\geq 3$ the set $K^n_{NSD}$ has
an atom, located within
the subspace $K^n_{NSD\vert\I^{(n-1)}}$ of non-selfdual algebras
with ideal $I^{(n-1)}$.
For $n\equiv 3\, \mbox{\rm mod}\, 4$ the atom is
nilpotent unimodular, given by $\ii(n)$,
and for $n\equiv 1\, \mbox{\rm mod}\, 4$ it is given by $\iv(n)$.
}
\hfill$\Box$ \break
The selfduality of the $K^n_{de}$-atoms $a_m(n)$ and $\ve(n)$
excludes them as candidates
for $K^n_{NSD}$-atoms.
It remains an open problem to determine
at least some $K^n_{NSD}$-atom for arbitrary even $n$,
and all $K^n_{NSD}$-atoms for arbitrary
$n$.
For odd $n$, besides $\ii(n)$ or $\iv(n)$, there might be further
$K^n_{NSD}$-atoms, even within
$K^n_{NSD\vert\I^{(n-1)}}$.
However, assume we succeed for some $n$ to determine all
$K^n_{NSD}$-atoms and
furthermore to show that $K^n_{NSD}$ is connected for that $n$.
In Sec. 7 we will actually see that, for $n=3$ the
{Heisenberg} algebra
$\ii(3)$ is the only non-selfdual atom, hence $K^3_{NSD}$ is connected,
and for $n=4$,
with the topology of $K^4$ obtained in Sec. 6.3
and the non-selfdual algebras of Sec. 7.2.2,
the resulting non-selfdual set $K^4_{NSD}$ will
be connected, and its explicit structure will
reveal the $K^4_{NSD}$-atoms.
Let us assume in the following that for a given $n$ the space
$K^n_{NSD}$ is connected.
For $n\equiv 3 \,\mbox{mod}\, 4$ resp. $n\equiv 1 \,\mbox{mod}\, 4$
corresponding to the $K^n_{NSD}$-atom $\ii(n)$ resp. $\iv(n)$ there are
in any case $2$ atoms of $K^n_{or,NSD}=K^n_+\oplus K^n_-$,
either $\ii(n)^R$ and $\ii(n)^L$, or
resp. $\iv(n)^R$ and $\iv(n)^L$.
Similarly, we could pick for arbitrary $n$ any $K^n_{NSD}$-atom $a$
and will find a corresponding pair of $K^n_{or,NSD}$-atoms $a^R$ and $a^L$.
At this place, let us make
the convention to assign the right atom $a^R$ to $K^n_+$ and
the left atom $a^L$ to $K^n_-$.
Now consider all other pairs of dual points $A^R$ and $A^L$ in
$K^n_{or,NSD}$, which constitute the preimage $\pi^{-1}(A)$ of a
non-selfdual point $A\in K^n_{NSD}$.
For any limit $A\to B$ or $C\to A$ in $K^n$ there
exists a corresponding pair of limits $A^{R/L}\to B'$ or $C'\to A^{R/L}$
in $K^n_{or}$, with $B'\in \pi^{-1}(B)$ resp. $C'\in \pi^{-1}(C)$.
Note however that there are no transitions or limits
between conjugate points,
neither $A^R\to A^L$ nor $A^L\to A^R$, because limits cannot reverse
the orientation.
If $B'$ or $C'$ is non-selfdual,
we demand it, as the limit $B'=B^{R/L}$ resp.
the prelimit $C'=C^{R/L}$ of $A^{R/L}$,
to be contained in the same component
of $K^n_{or,NSD}$ as $A^{R/L}$ itself.
Under consideration of the transitivity of
transitions in $K^n_{or}$ and use of the assumed connectedness
of $K^n_{or,NSD}$,
it follows from assignments for the non-selfdual
atoms made above that, {\em all} right algebras have to be in $K^n_+$ and
{\em all} left algebras have to be in $K^n_-$.
If $K^n_{NSD}$ is connected,
this choice is the only one which makes
each of $K^n_+$ and $K^n_-$ connected
and both disconnected to each other. Therefore
it is the canonical assignment in the case of connected
$K^n_{NSD}$. This will be
the relevant situation in the following sections.
{\hfill \break}
For $n\geq 4$ let us define a selfdual algebra $A^a_{n,2}$ with
{Abel}ian ideal $I^{(n-1)}$ by
$\NJNF(A^a_{n,2}):=[a\cdot\NJNF(A_2)] \oplus \NJNF(\iv(n-1))$, where $\oplus$
denotes the direct sum of matrices.
Now it is easy to prove
{\hfill \break}
{\bf Lemma 5.}
{\em
Within $K^n$ for $n\geq 3$, the subset $K^n_{SD}$ of selfdual elements in
$K^n_{(or)}$ has the following properties:
If there exists a non-selfdual algebra,
which is the case at least for $n$ odd, then $K^n_{SD}$ is not open.
For $n$ odd $K^n_{SD}$ is neither open nor closed.
} {\hfill \break}
{Proof:}
Assume that there exists a non-selfdual algebra;
such an algebra is given by $\iv(n)$ for $n$ odd.
Then there is at least one
$K^n_{NSD}$-atom.
Any $K^n_{NSD}$-atom has a selfdual limit.
Hence, there exists a
selfdual limit from a non-selfdual sequence
$ {\Rightarrow } K^n_{(or),NSD}$ not closed
$ {\Rightarrow } K^n_{SD}$ not open. {\hfill \break}
On the other hand, there are also non-selfdual limits from selfdual
sequences, like $\VIo\to\II$ for $n=3$
and $A^a_{n,2} {\longrightarrow } \iv(n)$ with $a\to 1$ for odd $n>3$
$ {\Rightarrow } K^n_{SD}$ not closed for odd $n\geq 3$.
\hfill$\Box$ \break
Likewise, each of $K^n_\pm$ is neither open nor closed for $n$ odd.
Note that $K^n_{SD}$ open would imply
$K^n_{SD}=K^n$.
In examination of duality of a given algebra, it is useful
to remind the obvious
{\hfill \break}
{\bf Lemma 6.}
{\em
For an algebra $A\in K^n$, following assertions are equivalent:
{\hfill \break}
i) $A$ is selfdual.
{\hfill \break}
ii) The set $S(A)$ of all subalgebras of $A$ is selfdual.
{\hfill \break}
iii) The set $J(A)$ of all ideals of $A$ is selfdual.
}
\hfill$\Box$ \break
Note that individual elements of $S(A)$ and $J(A)$ taken for themselves can
be non-selfdual while $A$ is selfdual.
Finally we deal with the case of simple {Lie} algebras.
{\hfill \break}
{\bf Lemma 7.}
{\em
Simple {Lie} algebras are not selfdual.
} {\hfill \break}
{Proof:}
A simple $n$-dimensional {Lie} algebra $A_n$ can
be characterized by a
{Cartan-Weyl} basis.
Such a basis consisting of generators
$H_i$, $i=1,\ldots,l=\mbox{rank}A_n$,
which span a maximal {Abel}ian subalgebra
(usually called {Cartan} subalgebra) and $n-l$ generators
$E_\alpha$, each satisfying,
for any nonvanishing generator
$H=\alpha^i H_i$ of the {Cartan} subalgebra,
a root equation
$[H, E_\alpha]=\alpha E_\alpha$ $\ (\ast)$
with root
$\alpha=\alpha^i\alpha_i$.
The commutators
$[E_\alpha,E_\beta]=N_{\alpha\beta}E_{\alpha+\beta}$ $\ (\ast\ast)$
for $\alpha+\beta\neq 0$ and
$[E_\alpha,E_{-\alpha}]=H$ $\ (\ast\ast\ast)$
are nonvanishing.
{}From the root equations $(\ast)$ we see that for any nonvanishing
{Cartan} subalgebra element $H$ (given by its coroots $\alpha^i$)
the reflection $H\to-H$ changes the algebra.
Furthermore by $(\ast\ast)$ and $(\ast\ast\ast)$ also none of the
reflections
$E_\alpha\to-E_\alpha$ keeps the
algebra invariant. Since there is no
reflection keeping
the algebra invariant it can not be selfdual.
\hfill$\Box$ \break
For considerations of the topological structure of $K^3$ and $K^4$
in Sec. 6 and 7 respectively, we will define the notion of parametrical
connectedness
of points in $K^n$
like following:
{\hfill \break}
{\bf Definition 7.}
$X, Y\in K^n$ are called {\em parametrically connected} iff
there exists a continuous curve
$c: [0,1] \to K^n$ with $c(0)=X$ and $c(1)=Y$ such that,
for all $t_1\leq t_2\in [0,1]$ with
$c(t_1)\neq c(t_2)$, there exist some $t_0\in [t_1,t_2]$
such that $c(t_1)\neq c(t_0)\neq c(t_2)$.
Otherwise $X, Y\in K^n$ are said to be
{\em parametrically disconnected}.
\hfill $\Box$\break
Note that, in the topology $\kappa^n$,
arcwise connectedness does not imply parametrical connectedness
as defined above.
Furthermore, a set $S\subset K^n$ is called parametrically connected,
iff any two points $X,Y\in S$ are parametrically connected in $S$.
$S\subset K^n$ is a {\em parametrically connected component}
iff $S$ is parametrically connected but not a proper subset of
another parametrically connected set.
{\hfill \break}
\section{\bf Topology of $K^n$ for $n\leq 4$}
\setcounter{equation}{0}
Sec. 6.1 resumes already existing results for $n\leq 3$,
Sec. 6.2 describes in detail the components and transitions
of $K^4$, and Sec. 6.3 gives some overview over the topological
structure of $K^4$.
\subsection{\bf Structure of $K^n$ for $n\leq 3$}
The {Lie} algebras with $n\leq 3$ are well known and listed, e.g.
by {Patera} and {Winternitz} \PaW.
$K^2$ contains only 2 elements, the {Abel}ian $2A_1$ and $A_2$ represented
by the algebra with $[e_2,e_1]=e_1$ as only nonvanishing bracket.
So $A_2$ has the ideal $J_1=A_1$ spanned by $\{e_1\}$, and is characterized
by $C_{<2>}=(1)\neq 0$ in contrast to $2A_1$. Note that
$A_2\equiv {\V}^{(2)}$. Obviously $\dim K^2_*=0$ and the unimodular
subset $U^2_*\subset K^2_*$ is empty.
The elements of $K^3$ correspond to the famous {Bianchi}
(or {Bianchi-Behr}) types.
They have been classified independently
first by S. {Lie} \Lie\ and then by L. {Bianchi} \Bi.
For their systematic derivation and explanation of their role for
cosmological models see e.g. {Kramer, Stephani} et al. \Kr.
For convenience of the reader we give
for each of the Bianchi types I up to IX an explicit description
by the commutators of its generators $e_1, e_2, e_3$ according
to {Landau-Lifschitz} \Lan:
Types I,II and VIII/IX are given by basic commutators
$$
[e_1,e_2]=n_3 e_3, [e_2,e_3]=n_1 e_1, [e_3,e_1]=n_2 e_2,
$$
with triplets $(n_1,n_2,n_3)$ respectively given by $(0,0,0), (1,0,0)$ and
$(1,1,\mp 1)$.
The 1-parameter families \VIh/ \mbox{${\rm VII}_h$} with $h\geq 0$ are given respectively
by
$$
[e_1,e_2]=e_3+he_2, [e_2,e_3]=0, [e_1,e_3]=\pm e_2+he_3,
$$
and especially it is $\III={\rm VI}_1$. IV resp. V are given by
$$
[e_1,e_2]=be_3+e_2, [e_2,e_3]=0, [e_1,e_3]=e_3
$$
with $b=1$ resp. $b=0$.
The 3-dimensional real {Lie} algebras in the notation of
{Patera} and {Winternitz} \PaW\ can be
characterized by their NJNF, which is
simultaneously the normal form
(see Eqs. (6.1) up to (6.4) below)
of the {Bianchi} types associated to them like in Table 1.
\bigskip {\hfill \break}
{\normalsize
\begin{tabular}{ccccccccccc}
$3A_1$&$A_1\oplus A_2$&$A_{3,1}$&$A_{3,2}$&$A_{3,3}$
&$A_{3,4}$&$A^a_{3,5}$&$A_{3,6}$&$A^a_{3,7}$&$A_{3,8}$&$A_{3,9}$\\
I&III&II&IV&V&\VIo&\VIh& \mbox{${\rm VII}_0$} & \mbox{${\rm VII}_h$} &VIII&IX
\end{tabular}
\begin{center}
\begin{tabular}{ll}
Table 1: &Inequivalent 3-dim. {Lie} algebras as denoted in \PaW\ (upper row)\\
&and corresponding {Bianchi} types (lower row).
\end{tabular}
\end{center}
\smallskip }
For convenience we explicitly give the nonvanishing commutators
for the indecomposable algebras $A_{3,1}$ up to $A_{3,9}$ from \PaW:
$$
A_{3,1}: [e_2,e_3]=e_1;
$$
$$
A_{3,2}: [e_1,e_3]=e_1, [e_2,e_3]=e_1+e_2;
$$
$$
A_{3,3}: [e_1,e_3]=e_1, [e_2,e_3]=e_2;
$$
$$
A_{3,4}: [e_1,e_3]=e_1, [e_2,e_3]=-e_2;
$$
$$
A^a_{3,5}: [e_1,e_3]=e_1, [e_2,e_3]=ae_2,\ 0<\vert a\vert<1;
$$
$$
A_{3,6}: [e_1,e_3]=-e_2, [e_2,e_3]=e_1;
$$
$$
A^a_{3,7}: [e_1,e_3]=ae_1-e_2, [e_2,e_3]=e_1+ae_2,\ 0<a;
$$
$$
A_{3,8}: [e_1,e_2]=e_1, [e_2,e_3]=e_3, [e_3,e_1]=2e_2;
$$
$$
A_{3,9}: [e_1,e_2]=e_3, [e_3,e_1]=e_2, [e_2,e_3]=e_1.
$$
In $3$ dimensions, all solvable algebras contain the {Abel}ian ideal
$J_2=2A_1$.
Therefore they can be characterized by their NJNF.
$$
\NJNF(\I) =
\left(
\begin{array}{cc}
0 & \\
& 0
\end{array}
\right) ,
$$
$$
\NJNF(\III) =
\left(
\begin{array}{cc}
1 & \\
& 0
\end{array}
\right) ,
$$
$$
\NJNF(\II) =
\left(
\begin{array}{cc}
0 & 1 \\
& 0
\end{array}
\right) ,
\qquad \II={\II}^{(3)} ,
$$
$$
\NJNF(\IV) =
\left(
\begin{array}{cc}
1 & 1 \\
& 1
\end{array}
\right) ,
\qquad \IV={\IV}^{(3)} ,
$$
\begin{equation}
\NJNF(\V) =
\left(
\begin{array}{cc}
1 & \\
& 1
\end{array}
\right) ,
\qquad \V={\V}^{(3)} .
\end{equation}
\begin{equation}
\NJNF(\VIo) =
\left(
\begin{array}{cc}
1 & \\
& -1
\end{array}
\right) ,
\
\NJNF(\VIh) =
\left(
\begin{array}{cc}
1 & \\
& a
\end{array}
\right) .
\end{equation}
In Eq. (6.2) the range $0< h< \infty, h\neq 1$ of the parameter
according to {Landau-Lifschitz} \Lan, denoted here by $h$,
is monotonously homeomorphic to the range
$-1< a< 1, a\neq 0$. $h=1$ resp. $a=0$ yields a decomposable algebra,
namely \III.
\begin{equation}
\NJNF( \mbox{${\rm VII}_0$} ) =
\left(
\begin{array}{cc}
0 & 1 \\
-1 & 0
\end{array}
\right) ,
\
\NJNF( \mbox{${\rm VII}_h$} ) =
\left(
\begin{array}{cc}
a & 1 \\
-1 & a
\end{array}
\right) .
\end{equation}
In Eq. (6.3) the range $0< h< \infty$ of the parameter
according to {Landau-Lifschitz} \Lan, denoted here by $h$,
is monotonously homeomorphic to the range
$0< a< \infty$. {\hfill \break}
Note that for a topological characterization of $K^3$ it is sufficient to
know the relation of the parameters $a$ and $h$ in (6.2) and (6.3)
at points of qualitative change in the NJNF and to ensure
homeomorphisms of the ranges between these critical points.
This is precisely the data we have given above. (Though the explicit
relation of $a$ and $h$ follows from the equivalence transform to normal form,
here we do not need to calculate it.)
Both \VIh and \mbox{${\rm VII}_h$} are unimodular for $h=0$ and converge to \II for
$0\leq h<\infty$ and to \IV for $h\to \infty$. {\hfill \break}
The simple algebras $ \mbox{${\rm VIII}$} =su(1,1)$ and $\IX=su(2)$
are described respectively by the 3 matrices
\begin{equation}
C_{<3>} =
\left(
\begin{array}{cc}
0 & 1 \\
-1 & 0
\end{array}
\right) ,
\qquad
C_{<1>} = -C_{<2>} =
\left(
\begin{array}{cc}
0 & 1 \\
\pm 1 & 0
\end{array}
\right) .
\end{equation}
$\NJNF(C_{<3>})=\NJNF( \mbox{${\rm VII}_0$} )$ for both $ \mbox{${\rm VIII}$} $ and $\IX$, {\hfill \break}
but
$\NJNF(C_{<1>})=\NJNF(C_{<2>})$ is equal to $\NJNF( \mbox{${\rm VII}_0$} )$ for $ \mbox{${\rm VIII}$} $ and to
$\NJNF(\VIo)$ for $\IX$. Therefore $\IX\to \mbox{${\rm VII}_0$} $, but $\IX\not\to\VIo$,
but both $ \mbox{${\rm VIII}$} \to\VIo$ and $ \mbox{${\rm VIII}$} \to \mbox{${\rm VII}_0$} $. {\hfill \break}
Considering all components, their parametrical limits and transitions
together,
we get the full topological structure of $K^3$,
which includes a transitive network of nearest neighbour
transitions between different components. The network has been depicted
already by {Mac Callum} \Mac\ and its transitivity
was outlined by {Schmidt} \aSch.
We have $\dim K^3=1$, since its largest parametrically connected components
are 1-dimensional. {\hfill \break}
For the unimodular subvariety $U^3_*\subset K^3_*$ it is
$\dim U^3_* =0$, and $\{ \mbox{${\rm VIII}$} ,\IX\}\subset U^3_*$ is a minimal dense
subset of isolated points.
Fig. 1 shows the transitive network of transitions in $K^3_*$,
with unimodular points encircled.
\vspace*{12.7truecm}
\begin{center}
{\normalsize Fig. 1: Transitive network of transitions in $K^3_*$.}
\end{center}
{\newpage } \noindent
\subsection
{\bf Components of $K^4$, transitions and parametrical limits}
The real 4-dimensional {Lie} algebras have been classified
by {Mubarakzyanov} \bMu\ and
listed by {Patera} and {Winternitz} \PaW. An early, somehow more
coarse classification has been given by {Petrov} \Pe.
For the convenience of the reader we explicitly give this classification
in terms of nonvanishing basic commutators. In order to avoid confusion with
the $3$-dimensional {Bianchi} types we alter the notation of {Petrov}'s
classes \Pe\ from I,\ldots,VIII to $\wp_i,\ i=1,\ldots,8$.
The subclasses ${\rm VI}_{1/4}$ are written as \PVIac respectively,
and ${\rm VI}_2$ together with ${\rm VI}_3$ are resumed in a single class
\PVIb in order to correspond to distinct classes of \PaW. With this
notation {Petrov}'s classes are characterized like following:
Solvable algebras, without {Abel}ian subgroup $3A_1$:
$$
\PI: [e_2,e_3]=e_1, [e_1,e_4]=ce_1, [e_2,e_4]=e_2, [e_3,e_4]=(c-1)e_3,\
c\in\R;
$$
$$
\PII: [e_2,e_3]=e_1, [e_1,e_4]=2e_1, [e_2,e_4]=e_2, [e_3,e_4]=e_2+e_3;
$$
$$
\PIII: [e_2,e_3]=e_1, [e_1,e_4]=qe_1, [e_2,e_4]=e_3, [e_3,e_4]=-e_2+qe_3,\
q^2<4;
$$
$$
\PIV: [e_2,e_3]=e_2, [e_1,e_4]=e_1;
$$
$$
\PV: [e_2,e_3]=e_2, [e_3,e_1]=-e_1, [e_1,e_4]=e_2, [e_2,e_4]=-e_1;
$$
Solvable algebras, with {Abel}ian subgroup $3A_1$:
$$
\PVIa: [e_1,e_4]=ae_1+be_4, [e_2,e_4]=ce_2+de_4, [e_3,e_4]=ee_3+fe_4,
$$
with real tuples $(a,b,c,d,e,f)$ of the form
$(0,0,0,0,0,0)$, $(0,1,0,1,0,0)$, $(0,1,0,1,0,1)$, $(1,1,0,0,0,0)$ or
$(1,0,c,0,e,0)$;
$$
\PVIb: [e_1,e_4]=ke_1+e_2, [e_2,e_4]=ke_2+de_3, [e_3,e_4]=ee_3,\
k\in \R,\ d,e\in\{0,1\};
$$
$$
\PVIc: [e_1,e_4]=ke_1+e_2, [e_2,e_4]=-e_1+ke_2, [e_3,e_4]=le_3,\
k,l\in \R;
$$
Non solvable algebras:
$$
\PVII: [e_1,e_2]=e_1, [e_2,e_3]=e_3, [e_3,e_1]=-2e_2;
$$
$$
\PVIII: [e_1,e_2]=e_3, [e_2,e_3]=e_1, [e_3,e_1]=e_2.
$$
{\newpage }
In the following we use the characterization of equivalence
classes by their NJNF, according to Sec. 3, in order to find relative
positions of the equivalence classes in $K^4$, possible
transitions between them and parametrical limits of parametrically
connected components of $K^4$.
{\hfill \break} \nl
{\bf 6.2.1 Decomposable {Lie} algebras}
{\hfill \break} \nl
A decomposable 4-dimensional {Lie} algebra can have the structures
$4A_1$, $2A_1\oplus A_2$, $2A_2$ or $A_1\oplus A_3$. The first 3
possibilities are unique, since $A_1$ is the unique 1-dim. {Lie} algebra
and $A_2$ is the unique non {Abel}ian 2-dim. {Lie} algebra. Note that
$2A_2\equiv \PIV$ in {Petrov}'s classification \Pe.
$A_1\oplus A_3$ consists of 9 classes, given by $\{A_{3,i}\}_{i=1,\ldots,9}$
listed
in Table 1. It is $A_1\oplus \II\equiv {\II}^{(4)}$.
$A_1\oplus \mbox{${\rm VIII}$} $
and $A_1\oplus\IX$ are the same as in \Pe\ the \PVII and \PVIII respectively.
{\hfill \break}
Transitions and limits: Besides the transitions and limits induced by
{Abel}ian embedding
$\oplus \R$ of transitions in $K^3$, there are further transitions,
which prevent the embedding $\oplus \R$ to be a homeomorphism.
So for example $\V\oplus \R\to \II\oplus \R$, but
$\V\not\to \II$.
This demonstrates that, while $\V$ is an atom, $\V\oplus \R$ is not.
Furthermore $\VIo\oplus \R$ and $ \mbox{${\rm VII}_0$} \oplus \R$ both go first to $A_{4,1}$
and then to $\II\oplus \R$.
$ \mbox{${\rm VIII}$} \oplus \R$ has a limit in the non decomposable
$A_{4,8}$ and, like $\IX\oplus \R$, also in $A_{4,10}$,
as described below.
The algebra $2A_2$ in spite of being decomposable is not the limit of
any other algebra in $K^4$. It has transitions to $\VIh\oplus \R$
with $h\geq 0$, to $A_{4,3}$ and to $A^0_{4,9}$.
{\hfill \break} \nl
\noindent
{\bf 6.2.2 Indecomposable {Lie} algebras}
{\hfill \break} \nl
Coarsely these algebras have already been classified by {Petrov} \Pe.
Table 2
relates his classification to that of {Patera} and
{Winternitz} \PaW.
\smallskip
{\normalsize
\begin{center}
\begin{tabular}{ccccccc}
$A_{4,1..4}$&$A_{4,5}$&$A_{4,6}$&$A_{4,7}$&$A_{4,8/9}$&$A_{4,10/11}$&$A_{4,12}$
\smallskip\\
\PVIb &\PVIa &\PVIc &\PII &\PI &\PIII &\PV
\end{tabular}\smallskip\\
\begin{tabular}{ll}
Table 2:&Classification of {Petrov} \Pe\ (lower row)
and \PaW\ (upper row)\\
& of 4-dim. {Lie} algebras except decomposable ones.
\end{tabular}
\end{center}
\smallskip }
The algebra \PIII, $q=0$ is the
same as $A_{4,10}$. It is the only indecomposable $4$-dimensional
algebra that corresponds to a maximal isometry group of a $3$-dimensional
homogeneous {Riemann}ian space
(see Sec. 8, 9 below and {Bona} and {Coll} \Bo, Theorem 1).
For convenience we explicitly give the nonvanishing commutators
of the indecomposable algebras $A_{4,1}$ up to $A_{4,12}$ according
to \PaW:
$$
A_{4,1}: [e_2,e_4]=e_1, [e_3,e_4]=e_2;
$$
$$
A^a_{4,2}: [e_1,e_4]=ae_1, [e_2,e_4]=e_2, [e_3,e_4]=e_2+e_3,\ a\neq 0;
$$
$$
A_{4,3}: [e_1,e_4]=e_1, [e_3,e_4]=e_2;
$$
$$
A_{4,4}: [e_1,e_4]=e_1, [e_2,e_4]=e_1+e_2, [e_3,e_4]=e_2+e_3;
$$
$$
A^{a,b}_{4,5}: [e_1,e_4]=e_1, [e_2,e_4]=ae_2, [e_3,e_4]=be_3,\
-1\leq a\leq b\leq 1,\ ab\neq 0;
$$
$$
A^{a,b}_{4,6}: [e_1,e_4]=ae_1, [e_2,e_4]=be_2-e_3, [e_3,e_4]=e_2+be_3,\
b\geq 0,\ a\neq 0;
$$
$$
A_{4,7}: [e_1,e_4]=2e_1, [e_2,e_4]=e_2, [e_3,e_4]=e_2+e_3, [e_2,e_3]=e_1;
$$
$$
A_{4,8}: [e_2,e_3]=e_1, [e_2,e_4]=e_2, [e_3,e_4]=-e_3;
$$
$$
A^b_{4,9}: [e_2,e_3]=e_1, [e_1,e_4]=(1+b)e_1, [e_2,e_4]=e_2, [e_3,e_4]=be_3,\
-1<b\leq 1;
$$
$A_{4,8}$ is the parametrical limit of $A^b_{4,9}$ for $b\to -1$;
hence by {Mubarakzyanov} \aMu\ and {Petrov} \Pe\
$A_{4,8}$ and $A_{4,9}$ are subsumed in a single
$1$-parameter set.
$$
A_{4,10}: [e_2,e_3]=e_1, [e_2,e_4]=-e_3, [e_3,e_4]=e_2;
$$
$$
A^a_{4,11}: [e_2,e_3]=e_1, [e_1,e_4]=2ae_1, [e_2,e_4]=ae_2-e_3,
[e_3,e_4]=e_2+ae_3,\ 0<a;
$$
$A_{4,10}$ is the parametrical limit of $A^a_{4,11}$ for $a\to 0$;
hence by {Mubarakzyanov} \aMu\ and {Petrov} \Pe\
$A_{4,10}$ and $A_{4,11}$ are subsumed in a single
$1$-parameter set.
$$
A_{4,12}: [e_1,e_3]=e_1, [e_2,e_3]=e_2, [e_1,e_4]=-e_2, [e_2,e_4]=e_1.
$$
The only difference of this classification to that of
{Mubarakzyanov} \aMu\ is that, unlike there, here
the endpoints $A_{4,8}$ and $A_{4,10}$ are distinguished
against the rest of the $1$-parameter sets $A_{4,9}$ and $A_{4,11}$
respectively.
In the following we reclassify these algebras by their NJNF.
{\hfill \break}
{\hfill \break}
{\bf a) Algebras with an {Abel}ian ideal $J_3=3A_1\equiv \I$}
{\hfill \break}
{\hfill \break}
These are the algebras of type \PVI. In the following the cases i) and ii)
correspond to \PVIb, case iii) to \PVIa and case iv) to
\PVIc.
\medskip\hfill\break
i) 1 eigenvalue with 1 Jordan block: {\hfill \break}
Either the eigenvalue is $ {\lambda } = 0$ or otherwise it can be normalized to
$ {\lambda } = 1$.
{\newpage }
\begin{equation}
\NJNF(A_{4,1}) =
\left(
\begin{array}{ccc}
0 & 1 & \\
& 0 & 1 \\
& & 0
\end{array}
\right) ,
\
\NJNF(A_{4,4}) =
\left(
\begin{array}{ccc}
1 & 1 & \\
& 1 & 1 \\
& & 1
\end{array}
\right) .
\end{equation}
Transitions: Obviously $A_{4,4}\to A_{4,1}$ and, by increasing the geometric
multiplicity, $A_{4,4}\to{\IV}^{(4)}\equiv A^1_{4,2}$ resp.
$A_{4,1}\to{\II}^{(4)}\equiv \II\oplus \R$.\medskip\hfill\break
ii) Maximally 2 eigenvalues with together 2 Jordan blocks: {\hfill \break}
Here JNF$(A)$ consists of both a $1\times 1$ and a $2\times 2$ Jordan block,
with eigenvalues $ {\lambda } _1$ and $ {\lambda } _2$ respectively. If $ {\lambda } _1 = 0$, the
algebra would become decomposable ($\II \oplus \R$ if $ {\lambda } _2 = 0$,
$\IV \oplus \R$ if $ {\lambda } _2 \neq 0$). Therefore assume $ {\lambda } _1 = a \neq 0$.
Either $ {\lambda } _2 = 0$, then $ {\lambda } _1 = 1$ after normalization, or
$ {\lambda } _2 \neq 0$, then it can be normalized to $ {\lambda } _2 = 1$. If
$ {\lambda } _2 = {\lambda } _1 = a$, there is only 1 eigenvalue, which can be
normalized to $a=1$. Note that $A^1_{4,2}\equiv {\IV}^{(4)}$, which is a
case to be considered separately.
\begin{equation}
\NJNF(A_{4,3}) =
\left(
\begin{array}{ccc}
1 & & \\
& 0 & 1 \\
& & 0
\end{array}
\right) ,
\
\NJNF(A^a_{4,2}) =
\left(
\begin{array}{ccc}
a & & \\
& 1 & 1 \\
& & 1
\end{array}
\right) .
\end{equation}
Transitions and limits: From $A^a_{4,2}$ with $a\neq 0,1$
to $\IV\oplus \R$ for $a\to 0$,
to $A_{4,4}$ for $a\to 1$, and to $A_{4,3}$ for $\vert a\vert\to\infty$. By
increasing the geometric multiplicity, to $A^{1,a}_{4,5}$ for $0<\vert a\vert
<1$,
to $A^{1,-1}_{4,5}=A^{-1,-1}_{4,5}$ for $a=-1$ and to
$A^{\frac{1}{a},\frac{1}{a}}_{4,5}$ for $1<\vert a\vert <\infty$.
Also generally $A^a_{4,2}\to A_{4,1}$. {\hfill \break}
{}From ${\IV}^{(4)}\equiv A^{1}_{4,2}$ to ${\II}^{(4)}\equiv \II \oplus \R$
and ${\V}^{(4)}\equiv A^{1,1}_{4,5}$, according to the remark at the theorem
in Sec. 4.
{\hfill \break}
{}From $A_{4,3}$ to $A_{4,1}$ and, by increasing of geometric multiplicity,
to $\III\oplus \R$.\medskip\hfill\break
\noindent
iii) 3 real eigenvalues as Jordan blocks: {\hfill \break}
Assuming the largest eigenvalue normalized to $ {\lambda } _1 = 1$, there remain
$ {\lambda } _2=a$ and $ {\lambda } _3=b$ with $-1\leq b\leq a\leq 1$. If $a\cdot b = 0$,
the algebra becomes decomposable ($a=b=0$ yields $\III \oplus \R$, for
$a=1,b=0$ it is $\V \oplus \R$, for $a=0,b=-1$ it is $\VIo \oplus \R$ and
otherwise $a=0$ or $b=0$ yields $\VIh \oplus \R$). Therefore assume
$a\cdot b\neq 0$. The case $a=b=1$ (single 3-fold degenerate eigenvalue)
corresponds to the pure vector type
$A^{1,1}_{4,5} \equiv \Vv \neq \V \oplus \R$.
In $A^{1,b}_{4,5}$ and $A^{a,a}_{4,5}$ there are 2 eigenvalues, one of them
2-fold degenerate.
{\newpage } \noindent
For the nondegenerate case, $-1\le b< a< 1$. Note that
$A^{a,b}_{4,5}=A^{b,a}_{4,5}$, since permutations are in ${\GL}(4)$.
\begin{equation}
\NJNF(A^{a,b}_{4,5}) =
\left(
\begin{array}{ccc}
1 & & \\
& a & \\
& & b
\end{array}
\right) .
\end{equation}
Transitions and limits: From $A^{a,b}_{4,5}, a>b,$
to $A^{\frac{1}{a}}_{4,2}$ for ${b\to a}$,
to $A^{b}_{4,2}$ for ${a\to 1}$.
To $A_{4,4}$ for ${{a\to 1}\atop{b\to 1}}$,
to $\IV\oplus \R$ for ${{a\to 1}\atop{b\to 0}}$,
to $ \mbox{${\rm VII}_h$} \oplus \R, 1<h<\infty,$ for ${{a\not\to 0,1}\atop{b\to 0}}$,
to $A_{4,3}$ for ${{a\to 0}\atop{b\to 0}}$,
to $\VIh\oplus \R, 0\leq h<1,$ for ${{a\to 0}\atop{b\not\to 0}}$,
and to $A^{-1}_{4,2}$ for ${b\to -1}$ and ${a\to \pm 1}$.
Also generally $A^{a,b}_{4,5}\to A_{4,1}$. {\hfill \break}
Note furthermore that $A^{a,-1}_{4,5}=A^{-a,-1}_{4,5}$. {\hfill \break}
{}From $A^{1,b}_{4,5}$
to ${\IV}^{(4)}\equiv A^{1}_{4,2}$ for ${b\to 1}$, and
to ${\V}\oplus \R$ for ${b\to 0}$.
Generally $A^{1,b}_{4,5}\to {\II}^{(4)}$. {\hfill \break}
{}From $A^{a,a}_{4,5}$
to ${\IV}^{(4)}\equiv A^{1}_{4,2}$ for ${a\to 1}$, and
to ${\III}\oplus \R$ for ${a\to 0}$.
Generally $A^{a,a}_{4,5}\to {\II}^{(4)}$. {\hfill \break}
{}From ${\V}^{(4)}\equiv A^{1,1}_{4,5}$ only to $4A_1$, according to the
theorem of Sec. 4.\medskip\hfill\break
iv) 1 real eigenvalue and 2 complex conjugates: {\hfill \break}
If $ {\lambda } _{2,3}=r(\cos\theta \pm i\sin\theta)$, by normalization
$r\sin\theta = 1$ can be achieved, if $ {\lambda } _2 \neq {\lambda } _3$ is
assured (otherwise the Jordan block becomes diagonal). Set then
$r\cos\theta = b$ and $ {\lambda } _1=a$. Demand $a\neq 0$ to exclude
decomposability ($a = 0$ yields $ \mbox{${\rm VII}_h$} \oplus \R$ and $b=0$ then
corresponds to $h=0$) and without restriction $b\geq 0$.
\begin{equation}
\NJNF(A^{a,b}_{4,6}) =
\left(
\begin{array}{ccc}
a & & \\
& b & 1 \\
&-1 & b
\end{array}
\right) .
\end{equation}
Transitions and limits: For $a\to 0$, $A^{a,b}_{4,6}\to \mbox{${\rm VII}_h$} \oplus \R$,
with $0\leq h<\infty$
corresponding to $0\leq b<\infty$.
For a fixed ratio $\frac{a}{b}$ and
$b\to\infty$ there is a limit to $A^{\frac{a}{b}}_{4,2}$, if $a\neq b$,
and to $A_{4,4}$, if $a=b$.
$A^{a,b}_{4,6}\to A_{4,3}$ for $b$ finite (esp. $b=0$) and
$\vert a\vert\to\infty$, and
$A^{a,b}_{4,6}\to \IV\oplus \R$ for $a$ finite (esp. $a=0$) and $b\to\infty$.
Also generally $A^{a,b}_{4,6}\to A_{4,1}$. {\hfill \break}
Note furthermore that $A^{a,0}_{4,6}=A^{-a,0}_{4,6}$.
{\hfill \break}
{\hfill \break}
{\newpage } \noindent
{\bf b) Algebras with a nilpotent ideal $J_3=A_{3,1}\equiv{\rm II}$}
{\hfill \break}
{\hfill \break}
In the following case i) corresponds to \PII, case ii) to \PI
and case iii) to \PIII.\medskip\hfill\break
i) 2 eigenvalues with together 2 Jordan blocks: {\hfill \break}
\begin{equation}
\NJNF(A_{4,7}) =
\left(
\begin{array}{ccc}
2 & & \\
& 1 & 1 \\
& & 1
\end{array}
\right) .
\end{equation}
Transitions: $A_{4,7}\to A^{2}_{4,2}$ for $J_3\to \I$.
Furthermore
$A_{4,7}\to A^{1}_{4,9}$.\medskip\hfill\break
ii) 3 real eigenvalues as Jordan blocks: {\hfill \break}
$$
\NJNF(A^{b}_{4,9}) =
\left(
\begin{array}{ccc}
1+b & & \\
& 1 & \\
& & b
\end{array}
\right) , 0 < \vert b\vert < 1 ,
$$
$$
\NJNF(A^{0}_{4,9}) =
\left(
\begin{array}{ccc}
1 & & \\
& 1 & \\
& & 0
\end{array}
\right) ,
\qquad
\NJNF(A^{1}_{4,9}) =
\left(
\begin{array}{ccc}
2 & & \\
& 1 & \\
& & 1
\end{array}
\right) ,
$$
\begin{equation}
\NJNF(A_{4,8}) =
\left(
\begin{array}{ccc}
0 & & \\
& 1 & \\
& &-1
\end{array}
\right) .
\end{equation}
Transitions: From $A^b_{4,9}$
to $A_{4,8}$ for $b\to -1$,
to $A^0_{4,9}$ for $b\to 0$, and
to $A_{4,7}$ for $b\to 1$.
Furthermore, for $J_3\to \I$, to
$A^{\frac{1}{1+b},\frac{b}{1+b}}_{4,5}$ if $0<b<1$, and to
$A^{{1+b},{b}}_{4,5}$ if $-1<b<0$. {\hfill \break}
For $J_3\to \I$, $A^1_{4,9}\to A^{\frac{1}{2},\frac{1}{2}}_{4,5}$
and $A_{4,8} \to \VIo\oplus \R$. $A^0_{4,9}$ goes to $\IV\oplus \R$
and further to $\V\oplus \R$.
Since $ \mbox{${\rm VIII}$} \oplus \R\to A_{4,8}$, the latter is a limit from
a decomposable algebra.\medskip\hfill\break
iii) 1 real eigenvalue and 2 complex conjugates: {\hfill \break}
$$
\NJNF(A^{a}_{4,11}) =
\left(
\begin{array}{ccc}
2a & & \\
& a & 1 \\
&-1 & a
\end{array}
\right) , a > 0 ,
$$
\begin{equation}
\NJNF(A_{4,10}) =
\left(
\begin{array}{ccc}
0 & & \\
& 0 & 1 \\
&-1 & 0
\end{array}
\right) .
\end{equation}
{\newpage }
Transitions: From $A^a_{4,11}$
to $A_{4,10}$ for $a\to 0$,
to $A^1_{4,9}$ for $a\to \infty$ and,
for $J_3\to \I$, to
$A^{{2a},{a}}_{4,6}$. {\hfill \break}
For $J_3\to \I$, $A_{4,10}\to \mbox{${\rm VII}_0$} \oplus \R$. Furthermore
both $ \mbox{${\rm VIII}$} \oplus \R\to A_{4,10}$ and $\IX\oplus \R\to A_{4,10}$.
{\hfill \break}
{\hfill \break}
{\bf c) Algebras with a pure vector type ideal $J_3=A_{3,3}\equiv{\rm V}$}
{\hfill \break}
{\hfill \break}
This case corresponds to type \PV.
\begin{equation}
\NJNF(A_{4,12}) =
\left(
\begin{array}{ccc}
0 & 1 & \\
-1& 0 & \\
& & 0
\end{array}
\right) .
\end{equation}
Transitions: $A_{4,12}$ goes to $\V\oplus \R$, to $ \mbox{${\rm VII}_h$} \oplus \R$,
especially for $J_3\to \I$ to $ \mbox{${\rm VII}_0$} \oplus \R$, and to $A_{4,9}$.
{\hfill \break}
{\hfill \break}
\subsection{\bf The topological structure of $K^4$}
Since we know all components, their parametrical limits and transitions
in $K^4$,
we can now put
them together, in order to determine the full topological structure
of $K^4$.
Fig. 2 a), b) and c)
show components of $K^4_*$, with $J_3$ equal to $\I$, $\II$ and $\V$
respectively, as parts of the transitive network of convergence. The dashed
lines in Fig. 2 a) indicate the $\kappa^4$ limit lines from lines
in Fig. 2 b).
$\dim K^4=2$, since its largest (parametrically connected) components are
2-dimensional.
{\hfill \break}
For the unimodular subvariety $U^4_*\subset K^4_*$ it is
$\dim U^4_* =1$. The union of
$ \mbox{${\rm VIII}$} \oplus \R$, $\IX\oplus \R$,
$\{A^{a,-a-1}_{4,5},-\frac{1}{2}<a<0\}$
and
$\{A^{-2b,b}_{4,6},0<b<\infty\}$ is a dense
subset of $ U^4_*$ and consists of a minimum number of parametrically
connected components, namely 2 isolated points and 2 isolated line segments.
In Fig. 2
the unimodular lines are dotted, and the unimodular points encircled.
{\newpage }
\vspace*{17.5truecm}
\begin{center}
{\normalsize Fig. 2 a: Transitions and limits at components of $K^4_*$ with
ideal \I.}
\end{center}
\vspace*{8.5truecm}
\begin{center}
{\normalsize Fig. 2 b: Transitions at components of $K^4_*$ with ideal \II.}
\end{center}
\vspace{6.5truecm}
\begin{center}
{\normalsize Fig. 2 c: Transitions at components of $K^4_*$ with ideal \V.}
\end{center}
\section{\bf Orientation duality in $K^n_{(or)}$ for $n\leq 4$}
\setcounter{equation}{0}
In this section we
examine in detail all points in $K^n_{or}$ for $n\leq 4$ under the
aspect of orientation duality.
In Sec. 7.1 the topological structure of $K^n_{or}$ for
$n\leq 3$ is analysed by use of the (O)NJNF, thus reproducing the
results listed by {Schmidt} \bSch.
The connected components $K^3_\pm$ are determined explicitly.
Using the same method, Sec. 7.2 analyses the orientation duality structure
of $K^4_{or}$ in detail. Especially we determine the connected components
$K^4_\pm$.
\subsection{\bf Structure of $K^n_{or}$ for $n\leq 3$}
The {Lie} algebras in $K^n$ for $n\leq 3$ have been classified in
Sec. 6.1
using their $n-1$-dimensional ideals and the NJNF.
Their orientation duality has already been listed by
{Schmidt} \bSch.
$K^2$ contains only 2 elements, the {Abel}ian $2A_1$ and $A_2$ represented
by the algebra with $[e_2,e_1]=e_1$ as only nonvanishing bracket. Both are
selfdual, because e.g. $e_1\to-e_1$ does not change the algebra.
So $K^2_{or}=K^2_{SD}=K^2$
The elements of $K^3$ correspond to the familiar Bianchi types.
In the following
we analyse the orientation duality by looking at the NJNF in $K^3$ and for
non-selfduality also considering the ONJNF, defining the elements of
$K^3_\pm$.
The solvable algebras in $K^3$ contain all the {Abel}ian ideal $J_2=2A_1$.
In Sec. 6.1 they are classified according to their NJNF.
Similarly the solvable algebras in $K^3_{or}$ can be classified
according to their ONJNF, which agrees the NJNF in the case of
selfduality. So the selfdual algebras in $K^3_{or}$ correspond to the
following cases of NJNF w.r.t. the {Abel}ian ideal $J_2$:
$$
\NJNF(\I) =
\left(
\begin{array}{cc}
0 & \\
& 0
\end{array}
\right) ,
\quad
\NJNF(\V) =
\left(
\begin{array}{cc}
1 & \\
& 1
\end{array}
\right) ,
$$
$$
\NJNF(\III) =
\left(
\begin{array}{cc}
1 & \\
& 0
\end{array}
\right) ,
$$
\begin{equation}
\NJNF(\VIo) =
\left(
\begin{array}{cc}
1 & \\
& -1
\end{array}
\right) ,
\
\NJNF(\VIh) =
\left(
\begin{array}{cc}
1 & \\
& a
\end{array}
\right) .
\end{equation}
The algebras $\I$ and $\III$ are selfdual, since they are
decomposable. All algebras in Eq. (7.1) invariant under the reflection
$e_1\to-e_1$, which guarantees their selfduality.
The parameter range $0< h< \infty, h\neq 1$
($h$ denoting the parameter of {Landau-Lifschitz} \Lan),
corresponds monotonously to
$-1< a< 1, a\neq 0$.
$h=1$ resp. $a=0$ yields the decomposable \III.
So $K^3_{SD}=\{\I,\V\}\cup\{\VIh,0\leq h<\infty\}$.
The other solvable algebras which are not invariant under any reflection
are non-selfdual.
According to Sec. 3 and Sec. 4 we choose the reflection $e_3\to-e_3$
to characterize them as algebras in $K^3_\pm$, with their ONJNF
respectively given like following:
$$
\mbox{${\rm ONJNF}$} \{\IIRL\}=\pm\NJNF(\II) =
\pm
\left(
\begin{array}{cc}
0 & 1 \\
& 0
\end{array}
\right) ,
$$
$$
\mbox{${\rm ONJNF}$} \{\IVRL\}=\pm\NJNF(\IV) =
\pm
\left(
\begin{array}{cc}
1 & 1 \\
& 1
\end{array}
\right) ,
$$
$$
\mbox{${\rm ONJNF}$} \{ \mbox{${\rm VII}^{R/L}_0 $} \}=\pm\NJNF( \mbox{${\rm VII}_0$} ) =
\pm\left(
\begin{array}{cc}
0 & 1 \\
-1 & 0
\end{array}
\right) ,
$$
\begin{equation}
\mbox{${\rm ONJNF}$} \{ \mbox{${\rm VII}^{R/L}_h $} \}=\pm\NJNF( \mbox{${\rm VII}_h$} ) =
\pm\left(
\begin{array}{cc}
a & 1 \\
-1 & a
\end{array}
\right) .
\end{equation}
In Eq. (7.2) the parameter range $0< h< \infty$
($h$ denoting the parameter of {Landau-Lifschitz} \Lan)
corresponds monotonously to the range
$0< a< \infty$.
The simple algebras $ \mbox{${\rm VIII}$} =su(1,1)$ and $\IX=su(2)$
are described respectively by the 3 matrices
$$
C_{<3>} =
\left(
\begin{array}{cc}
0 & 1 \\
-1 & 0
\end{array}
\right) ,
\qquad
C_{<1>} = -C_{<2>} =
\left(
\begin{array}{cc}
0 & 1 \\
\pm 1 & 0
\end{array}
\right) .
$$
$\NJNF(C_{<3>})=\NJNF( \mbox{${\rm VII}_0$} )$ for both $ \mbox{${\rm VIII}$} $ and $\IX$, {\hfill \break}
but
$\NJNF(C_{<1>})=\NJNF(C_{<2>})$ is equal to $\NJNF(\VIo)$ for $ \mbox{${\rm VIII}$} $
and to
$\NJNF( \mbox{${\rm VII}_0$} )$ for $\IX$.
In the {Cartan-Weyl} basis $H:=-ie_3$, $E_{\pm}:=e_1\pm i e_2$
the nonvanishing commutators are given
as $[H,E_\pm]=\pm E_\pm$ and, for \mbox{${\rm VIII}$} or \IX respectively,
$[E_+,E_-]=\pm 2H$.
Note that the latter are different real sections in the same complex algebra.
According to Lemma 4.5 neither $ \mbox{${\rm VIII}$} $ nor $\IX$ are selfdual.
We discriminate the right and left algebra by the reflection
$e_3\to -e_3$,
defining both pairs $ \mbox{${\rm VIII}^{R/L} $} $ and $\IXRL$ of points in
$K^3_\pm$. So it is
\begin{equation}
\mbox{${\rm ONJNF}$} \{C^{R/L}_{<3>}\}= \mbox{${\rm ONJNF}$} \{ \mbox{${\rm VII}^{R/L}_0 $} \}
\end{equation}
and $C_{<1>}$ and $C_{<2>}$ interchange under this reflection.
{\hfill \break}
The table below summarizes the duality properties of the Bianchi classes in
$K^3$. Note that for any point $A \in K^3 {\setminus } K_{SD}$ there exists a
pair $(A^R,A^L) \in K_+\oplus K_-$ of points in $K^3_{or} {\setminus } K_{SD}$
with right/left handed bases respectively.
\medskip {\hfill \break}
{\normalsize
\begin{tabular}{ccccccccccc}
$3A_1$&$A_1\oplus A_2$&$A_{3,1}$&$A_{3,2}$&$A_{3,3}$
&$A_{3,4}$&$A^a_{3,5}$&$A_{3,6}$&$A^a_{3,7}$&$A_{3,8}$&$A_{3,9}$\\
I&III&II&IV&V&\VIo&\VIh& \mbox{${\rm VII}_0$} & \mbox{${\rm VII}_h$} &VIII&IX\\
1&1 &0 &0 &1&1 &1 &0 &0 &0 &0
\end{tabular}
\begin{center}
\begin{tabular}{ll}
Table 1: &3-dimensional {Lie} algebra classes in $K^3$,\\
&corresponding Bianchi types and selfduality (yes=1/no=0)
\end{tabular}
\end{center}
\medskip }
The non-selfdual subset of $K^3_{or}$ has 2 connected $1$-dimensional
components, $K^3_+$ and $K^3_-$, given respectively by
\begin{equation}
\mbox{${\rm VIII}^{R/L} $} /\IXRL\to \mbox{${\rm VII}^{R/L}_0 $} \gets \mbox{${\rm VII}^{R/L}_h $} \to\IVRL\to\IIRL ,
\end{equation}
where $\IIRL$ is respectively the atom of $K^3_\pm$.
\subsection{\bf Structure of $K^4_{or}$}
In Sec. 6.2 we classified the real 4-dimensional {Lie} algebras.
In this section they are reconsidered under the aspect
of orientation duality.
{\hfill \break} \nl
{\bf 7.2.1 Selfdual {Lie} algebras}
{\hfill \break} \nl
There exist following types of selfdual algebras:
a) all decomposable ones, b) indecomposable ones with ideal \I,
and c) some indecomposable ones with ideal \II.
{\hfill \break}
{\hfill \break}
{\bf a) Decomposable ones:}
{\hfill \break}
{\hfill \break}
All decomposable {Lie} algebras are selfdual.
A decomposable 4-dimensional {Lie} algebra can have the structures
$4A_1$, $2A_1\oplus A_2$, $2A_2$ or $A_1\oplus A_3$. The first 3
possibilities are unique, since $A_1$ is the unique 1-dim. {Lie} algebra
and $A_2$ is the unique non{Abel}ian 2-dim. {Lie} algebra.
$A_1\oplus A_3$ consists of 9 classes, given by $\{A_{3,i}\}_{i=1,\ldots,9}$
listed in Table 1. Note that $A_1\oplus \II\equiv {\II}^{(4)}$.
{\hfill \break}
{\hfill \break}
{\newpage } \noindent
{\bf b) Indecomposable ones with
ideal $J_3={\rm I}$:}
{\hfill \break}
{\hfill \break}
Algebras with
{Abel}ian ideal $J_3=3A_1\equiv \I$ are selfdual.
They
are given by the following cases:
{\hfill \break}
i) 1 Jordan block:
{\hfill \break}
These algebras are invariant under a combination
of the $3$ reflections $e_i\to -e_i$, $i=1,\ldots,3$.
Either the eigenvalue is $ {\lambda } = 0$ or otherwise it can be normalized to
$ {\lambda } = 1$.
$$
\NJNF(A_{4,1}) =
\left(
\begin{array}{ccc}
0 & 1 & \\
& 0 & 1 \\
& & 0
\end{array}
\right) ,
$$
\begin{equation}
\NJNF(A_{4,4}) =
\left(
\begin{array}{ccc}
1 & 1 & \\
& 1 & 1 \\
& & 1
\end{array}
\right) .
\end{equation}
These algebras are a $4$-dimensional analogue to \II and \IV.
While the latter are non-selfdual their even dimensional analogues
are selfdual.
These algebras are the essential dimensional ones, introduced
in Sec. 5 and denoted by $\ii(4)$ and $\iv(4)$.
$\ii(4)$ is an essential dimensional atom.
{\hfill \break}
{\hfill \break}
ii) 2 Jordan blocks:
{\hfill \break}
All these algebras are all invariant under the reflection $e_1\to -e_1$.
\begin{equation}
\NJNF(A_{4,3}) =
\left(
\begin{array}{ccc}
1 & & \\
& 0 & 1 \\
& & 0
\end{array}
\right) ,
\
\NJNF(A^a_{4,2}) =
\left(
\begin{array}{ccc}
a & & \\
& 1 & 1 \\
& & 1
\end{array}
\right) .
\end{equation}
In the latter case $a\neq 0$ and $A^1_{4,2}\equiv \IV^{(4)}$.
{\hfill \break}
{\hfill \break}
iii) 3 real eigenvalues as Jordan blocks:
{\hfill \break}
All these algebras are all invariant under the reflection $e_1\to -e_1$.
Assuming the largest eigenvalue normalized to $ {\lambda } _1 = 1$, there remain
$ {\lambda } _2=a$ and $ {\lambda } _3=b$ with $-1\leq b\leq a\leq 1$. If $a\cdot b = 0$,
the algebra becomes decomposable ($a=b=0$ yields $\III \oplus \R$, for
$a=1,b=0$ it is $\V \oplus \R$, for $a=0,b=-1$ it is $\VIo \oplus \R$ and
otherwise $a=0$ or $b=0$ yields $\VIh \oplus \R$). Therefore assume
$a\cdot b\neq 0$. The case $a=b=1$ (single 3-fold degenerate eigenvalue)
corresponds to the pure vector type
$A^{1,1}_{4,5} \equiv \Vv \neq \V \oplus \R$.
In $A^{1,b}_{4,5}$ and $A^{a,a}_{4,5}$ there are 2 eigenvalues, one of them
2-fold degenerate. For the nondegenerate case, $-1\le b< a< 1$. Note that
$A^{a,b}_{4,5}=A^{b,a}_{4,5}$, since permutations are in ${\GL}(4)$.
\begin{equation}
\NJNF(A^{a,b}_{4,5}) =
\left(
\begin{array}{ccc}
1 & & \\
& a & \\
& & b
\end{array}
\right) .
\end{equation}
\noindent
iv) 1 real eigenvalue and 2 complex conjugates:
{\hfill \break}
All these algebras are all invariant under the reflection $e_1\to -e_1$.
If $ {\lambda } _{2,3}=r(\cos\theta \pm i\sin\theta)$, by normalization
$r\sin\theta = 1$ can be achieved, if $ {\lambda } _2 \neq {\lambda } _3$ is
assured (otherwise the Jordan block becomes diagonal). Set then
$r\cos\theta = b$ and $ {\lambda } _1=a$. Demand $a\neq 0$ to exclude
decomposability ($a = 0$ yields $ \mbox{${\rm VII}_h$} \oplus \R$ and $b=0$ then
corresponds to $h=0$) and without restriction $b\geq 0$.
\begin{equation}
\NJNF(A^{a,b}_{4,6}) =
\left(
\begin{array}{ccc}
a & & \\
& b & 1 \\
&-1 & b
\end{array}
\right) .
\end{equation}
{\hfill \break}
{\hfill \break}
{\bf c) Indecomposable ones with ideal $J_3={\rm II}$:}
{\hfill \break}
{\hfill \break}
There exist algebras with non-selfdual ideal \II, which
are selfdual.
\begin{equation}
\mbox{${\rm ONJNF}$} (A_{4,8})=\NJNF(A_{4,8}) =
\left(
\begin{array}{ccc}
0 & & \\
& 1 & \\
& &-1
\end{array}
\right).
\end{equation}
This algebra is left invariant by a combination of reflections
$e_4\to-e_4$, $e_1\to-e_1$ and $e_2\leftrightarrow e_3$.
\begin{equation}
\mbox{${\rm ONJNF}$} (A_{4,10})=\NJNF(A_{4,10}) =
\pm\left(
\begin{array}{ccc}
0 & & \\
& 0 & 1 \\
&-1 & 0
\end{array}
\right) .
\end{equation}
This algebra is left invariant by a combination of reflections
$e_4\to-e_4$, $e_1\to-e_1$ and $e_2\to -e_2$.
{\newpage } \noindent
{\bf 7.2.2 Non-selfdual {Lie} algebras}
{\hfill \break} \nl
This kind of algebras exists with a basic ideal $J_3$, given either by the
non-selfdual \II or by the selfdual \V.
For all of them we have dual pairs of right and left points in $K^n_{or}$,
which transform to each other by $e_4\to -e_4$, constituting
by Sec. 3 and Sec. 4 the connected components $K^4_\pm$ respectively.
{\hfill \break}
{\hfill \break}
{\bf a) Indecomposable ones with ideal $J_3={\rm II}$:}
{\hfill \break}
{\hfill \break}
The ideal \II is non-selfdual.
For an algebra $A$ of the kinds listed below the there
exists no reflection leaving the set $J(A)$
invariant.
$$
\mbox{${\rm ONJNF}$} (A^{R/L}_{4,7})=\pm\NJNF(A_{4,7}) =
\pm\left(
\begin{array}{ccc}
2 & & \\
& 1 & 1 \\
& & 1
\end{array}
\right) ,
$$
$$
\mbox{${\rm ONJNF}$} (A^{b,R/L}_{4,9})=\pm\NJNF(A^{b}_{4,9}) =
\pm\left(
\begin{array}{ccc}
1+b & & \\
& 1 & \\
& & b
\end{array}
\right) , 0 < \vert b\vert < 1 ,
$$
$$
\mbox{${\rm ONJNF}$} (A^{0,R/L}_{4,9})=\pm\NJNF(A^{0}_{4,9}) =
\pm\left(
\begin{array}{ccc}
1 & & \\
& 1 & \\
& & 0
\end{array}
\right) ,
$$
$$
\mbox{${\rm ONJNF}$} (A^{1,R/L}_{4,9})=\pm\NJNF(A^{1}_{4,9}) =
\pm\left(
\begin{array}{ccc}
2 & & \\
& 1 & \\
& & 1
\end{array}
\right) ,
$$
\begin{equation}
\mbox{${\rm ONJNF}$} (A^{a,R/L}_{4,11})=\pm\NJNF(A^{a}_{4,11}) =
\pm\left(
\begin{array}{ccc}
2a & & \\
& a & 1 \\
&-1 & a
\end{array}
\right) , a > 0 .
\end{equation}
{\hfill \break}
{\hfill \break}
{\bf b) Indecomposable ones with ideal $J_3={\rm V}$:}
{\hfill \break}
{\hfill \break}
The only case here is given by
\begin{equation}
\mbox{${\rm ONJNF}$} (A^{R/L}_{4,12})=\pm\NJNF(A_{4,12}) =
\pm\left(
\begin{array}{ccc}
0 & 1 & \\
-1& 0 & \\
& & 0
\end{array}
\right) .
\end{equation}
Note that besides the selfdual ideal \V there is a second ideal \mbox{${\rm VII}_0$}
which is not selfdual, causing here the subset of ideals
$S(A_{4,12})$ to be non-selfdual.
Hence $A_{4,12}$ itself is non-selfdual.
{\hfill \break}
{\hfill \break}
{\bf c) The space $K^4_{NSD}$ and its components $K^4_\pm$:}
{\hfill \break}
{\hfill \break}
Collecting the algebras of the previous subsections a) and b) and recalling
transitions and parametrical limits of components
in $K^4_{NSD}$ according to Sec. 6, we find that
$K^4_{NSD}$ is connected, and so is each of $K^4_\pm$.
There are $2$ pairs of $K^4_{or,NSD}$-atoms, given by
$A^{0,R/L}_{4,9}$ and $A^{1,R/L}_{4,9}$.
In $n=4$ all $K^4_{NSD}$-atoms have an ideal $\II$, and hence
in the complement of the subspace $K^4_{NSD\vert\I}$ of
$K^4_{NSD}$-algebras with ideal $\I$.
Let us assign e.g. $A^{0,R/L}_{4,9}$ to $K^4_{\pm}$
respectively. Then the connectedness of $K^4_{\pm}$ and
the orientation preservation of limits within $K^4_{or,NSD}$
imply the assignment $A^{R/L}$ to $K^4_{\pm}$ respectively.
Note that with these assignments
the component $K^4_+$ is given as
\begin{equation}
\begin{array}{lcrr}
&A^{R}_{4,12}& \\
&\downarrow& \\
A^{-1<b<0,R}_{4,9}\to\!&A^{0,R}_{4,9}&\!
\gets A^{0<b<1,R}_{4,9}\to A^{R}_{4,7}\to A^{1,R}_{4,9}\gets A^{a>0,R}_{4,11}
\end{array}
\end{equation}
and the component $K^4_-$ as
\begin{equation}
\begin{array}{lcrr}
&A^{L}_{4,12}& \\
&\downarrow& \\
A^{-1<b<0,L}_{4,9}\to\!&A^{0,L}_{4,9}&\!
\gets A^{0<b<1,L}_{4,9}\to A^{L}_{4,7}\to A^{1,L}_{4,9}
\gets A^{a>0,L}_{4,11}
\end{array}
\end{equation}
So the non-selfdual components of $K^4_{(or)}$ are $1$-dimensional.
Note that $A^{R/L}_{4,1}\equiv \ii(4)^{R/L}$ is the atom of $K^4_\pm$
respectively.
\section{\bf Discussion and outlook}
\setcounter{equation}{0}
In Sec. 6 we determined {Lie} algebra transitions
in $K^4$ as limits induced by the topology $\kappa^4$.
Any {In\"on\"u-Wigner} contraction corresponds
to a certain transition; explicitly any of the
{In\"on\"u-Wigner} contractions listed in the tables of
{Huddleston} \Hud\ for real $4$-dimensional {Lie} algebras
corresponds to a transition in $K^4$.
Since {In\"on\"u-Wigner} contractions are only a special
case of the more general {Saletan} contractions,
and since even the latter do not induce all possible transitions
in $K^n$ with $n\geq 3$, it should not be surprising that we
have obtained transitions, which do not correspond to any
{In\"on\"u-Wigner} contraction, like e.g.
transitions $\IX\oplus \R\to A_{4,10}$, $ \mbox{${\rm VIII}$} \oplus \R\to A_{4,10}$
and transitions from $A^{a,b}_{4,5}$, $A^{a,b}_{4,6}$, $\VIh\oplus \R$
and $ \mbox{${\rm VII}_h$} \oplus \R$ to $A_{4,1}$.
The transition $\IX\oplus \R\to A_{4,10}$ corresponds to a
{Lie} algebra contraction, which was given already in \Sal\
(see Eqs. (35') to (37)) as an example of a {Saletan}
contraction, which can not be obtained as
a {In\"on\"u-Wigner} contraction.
It is also interesting to consider transitions in $K^3$ as obtained
in Sec. 5. The {In\"on\"u-Wigner}
contractions for real {Lie} algebras of dimension $d\leq 3$
are classified already by {Conatser} \Co.
The sequence of transitions
$ \mbox{${\rm VIII}$} \to\VIo\to\II\to\I$ is generated
by an iterated {Saletan} contraction (see \Sal, Eqs. (30) and (31)),
applied first to the
{Lie} algebra $ \mbox{${\rm VIII}$} $, of the $3$-dimensional
homogenous {Lorentz} group. On the 4-point subset
$\{ \mbox{${\rm VIII}$} ,\VIo,\II,\I\}$ {Saletan} contractions are transitive.
However this transitivity does not hold for {Saletan} transitions
on general subsets of $K^3$.
The sequence of transitions
$\IX\to \mbox{${\rm VII}_0$} \to\II\to\I$,
starting from the {Lie} algebra $\IX$ of the 3-dimensional
{Euclid}ean group, can not be obtained
by {Saletan} contractions.
The only {Saletan} contractions starting from $\IX$ are in fact
given by a {In\"on\"u-Wigner} contraction
$\IX\to \mbox{${\rm VII}_0$} $ and the trivial contraction
$\IX\to \I$. Though there exists a different
{In\"on\"u-Wigner} contraction
corresponding to the transitions $ \mbox{${\rm VII}_0$} \to\II$ there is no
{Saletan} contraction corresponding to $\IX\to\II$
(for a proof see \Sal). This example shows that, on an arbitrary
subset of $K^n$ with $n\geq 3$, in general not every transition
can be obtained from a {Saletan} contraction.
It implies that, even
on a set of points connected by {In\"on\"u-Wigner} contractions,
neither {Saletan}
contractions nor {In\"on\"u-Wigner} contractions need to be transitive.
Since we consider transitions between different points in $K^n$,
improper contractions of an algebra to an equivalent one can not be
seen by our method. For $n=4$ {Huddleston} \Hud\ identified
two types of algebras which admit only trivial and improper contractions.
These are precisely the two atoms of $K^4$, namely the unimodular
${\II}^{(4)}\equiv\II\oplus \R$ and
the pure vector type ${\V}^{(4)}\equiv A^{1,1}_{4,5}$.
For arbitrary dimension $n$, the atoms
of $K^n$ have been introduced and described first by
{Schmidt} {\bSch}.
By now the topological properties of $K^n$ for $n\leq 4$ have been examined.
It is natural to demand an investigation for arbitrary dimension $n$.
Practically, this is obstructed by the rapidly increasing number of
equivalence classes for increasing $n$.
A classification
for all nilpotent algebras has been done for $n=6$ by {Morozov} \Mo\
and for $n=7$ by {Ancochea-Bermudez} and {Goze} \An\ in the
complex case and by {Romdhani} \Ro, who
distinguishes $132$ components of real indecomposable nilpotent
$7$-dimensional {Lie} algebras.
For $K^5$ a full classification of all
real {Lie} algebras still distinguishes $40$ components
(compare {Mubarakzjanov} \bMu\ and {Patera} et al. \PaSWZ).
A determination of all possible transitions would be a rather tidy work.
However it is known by {\PaSWZ}
that $\dim K^5=3$,
because the maximal dimension of its components is 3.
Unlike for the classification of subalgebra
structures of each class in
$K^n$ (see {Patera, Winternitz} \PaW, and {Grigore, Popp} \Gri),
for the determination
of all equivalence classes and transitions between them there exists
no algorithm at present. However, a systematic exploitation of the NJNF,
which has been defined for arbitrary $n$, may contribute
some part to further progress.
The NJNF has proven to be a useful tool in characterizing
distinct $n$-dimensional {Lie} algebras with a common ideal
$J_{n-1}$ as endomorphisms ad$e_n$ of a complementary generator
$e_n$ on that ideal, with characteristic Jordan blocks
of their eigenvalues normalized by an overall scale.
In $4$ dimensions, besides decomposable
algebras, only cases with ideal \I, \II, or \V appear.
In $4$ dimensions, there are no simple algebras.
In general for $n\geq 6$
further classes of simple {Lie} algebras arise, which
lead to an additional further sophistication, as compared to $n=3$.
A combination of the established knowledge on semisimple {Lie} algebras
with the full classification of all {Lie} algebras would be desirable,
but is practically far away, since the dimensionality of the simple
{Lie} algebras increases rapidly with their rank. Note that all
simple components belong to the unimodular subset $U^n_*$.
In Sec. 7 we found that simple {Lie} algebras are non-selfdual
w.r.t. orientation reflection.
In general, we have neither a formula for $\dim K^n$ nor for
$\dim U^n_*$.
The $T_0$ topology allows components of different dimensions
to converge pointwise to each other, i.e. such that any point of the first
component converges to some point of the second component and any point of
the second component is the limit of some point of the first.
We have determined the topology of the space $K^n_{or}$ for $n\leq 4$.
The essential difference to $K^n$ is that the single non-selfdual
component of the latter is doubled to two components $K^n_+$
and $K^n_-$.
For $n=3$ or $4$, the space $K^n_{NSD}$ is nonvoid and connected.
For $n=3$ there is a unique $K^3_{NSD}$-atom
$\ii(3)=\II$.
$K^4_{NSD}$ has two atoms, $A^0_{4,9}$ and $A^1_{4,9}$,
and $A^{0<b<1}_{4,9}$ has boundary limits to both of them.
If $K^n_{NSD}$ is connected,
the arbitrariness in assigning conjugate pairs of points to $K^n_\pm$
can be reduced to a single decision for one pair only,
if we demand that both of $K^n_\pm$ are connected and to each other
disconnected.
At present, for general $n\geq 5$ it is not known whether
$K_{NSD}$ is connected.
{}From Eqs. (7.4) and (7.13-14), we see that the non-selfdual subset
$K^n_{NSD(,or)}$ of $K^n_{(or)}$ is $1$-dimensional
for both, $n=3$ and $n=4$.
We have $\dim K^3_{(or)}=1$ and $\dim K^n_{(or)}\geq 2$ for
dimension $n\geq 4$: In the latter case the
contribution of the non-selfdual subset is
of dimension less than that of the highest-dimensional
component,
while in the former case it is of highest dimension. Actually, the topology
of the highest-dimensional component of $K^3_{or}$ differs
from that of $K^3$ essentially.
The question of dimensionality for general $n\geq 5$ remains open,
for the non-selfdual subset as well as for $K^n_{(or)}$ itself.
Partial progress has been made by determining a candidate of an atom
of the non-selfdual
subset in odd dimension $n$. We found an interesting
periodicity in the structure of this $K^n_{NSD}$-atom: it is
$\ii(n)$ for $n=3 \,\mbox{mod}\, 4$
and $\iv(n)$ for $n=1 \,\mbox{mod}\, 4$.
However it remains an open problem to determine all atoms of
$K^n_{NSD}$ for arbitrary $n\geq 5$. Presently we do not know
how an atom for even $n\geq 6$ looks like in general.
With Definition 6 the notion of an atom from
{Schmidt} \bSch\ has been generalized to arbitrary subsets.
Although the present work is on the case of real {Lie} algebras,
we want to make some comments on the analogous complex cases
to the pairs of algebras $ \mbox{${\rm VII}_h$} /\VIh$ ($h\geq 0$), $\IX/ \mbox{${\rm VIII}$} $ of $K^3$ and
$A^{a,b}_{4,6}/A^{a,b}_{4,5}$, $A^{a}_{4,11}/A^{b}_{4,9}$ and
$A_{4,10}/A_{4,8}$ of $K^4$.
If one considers the analogous $3$- or $4$-dimensional
{Lie} algebras over the
complex basic field the group $\GL(n)$ is now correspondingly the
group of nonsingular complex linear transformation.
The pairs of complex conjugated eigenvalues associated to the
$2\times 2$ Jordan block of each of the first algebras of the pairs above
in the complex remain as $2$ Jordan blocks in a corresponding
complex JNF. After introducing a similar normalization convention
like for the real case, the complex analogues of the
NJNF will be the same for members of any pair above.
(For $n=3$ this had already been realized by
{Bianchi} \Bi. The complex $4$-dimensional case was considered
already by {Lie} \Lie).
{\hfill \break} \nl
{\Large {\bf Acknowledgments}}
{\hfill \break} \nl
I would like to express my gratitude to H.-J.
{ Schmidt} for
valuable discussions on the present topic.
{\newpage } \noindent
{\Large {\bf References}}
{\hfill \break} \nl
\An\ { J. M. Ancochea-Bermudez} and { M. Goze}.
Classification des algebres de { Lie} nilpotentes de dimension 7.
Arch. Math. {\bf 52}, 175-185 (1989).
\smallskip
{\hfill \break}
-, Sur la classification des algebres de { Lie} nilpotentes de dimension 7.
Comptes Rendus Acad. Sci. Paris, Ser. I {\bf 302}, 611-613 (1986).
\smallskip
{\hfill \break}
\Bi\ { L. Bianchi},
Sugli spazii a tre dimensioni che ammettono un gruppo continuo di
movimenti.
Soc. Ital. Sci. Mem. Mat., Ser. {IIIa}, {\bf 11}, 267 (1897).
\smallskip
{\hfill \break}
\bBi\ -,
Lezioni sulla teoria dei gruppi continui finiti di trasformazioni,
Cap. XIII: Applicazioni alla teoria degli spazii pluridimensionali
con un gruppo continuo di movimenti,
\S\S 198, 199 (for a description of types \I-\VII and \mbox{${\rm VIII}$} /\IX)
and \S\S 200-207 (for applications to homogeneous spaces), p. 550-578.
Spoerri, Pisa 1918.
\smallskip
{\hfill \break}
\Cha\ { R. Charles},
Sur la structure des algebres de { Lie} rigides.
Ann. Inst. Fourier Grenoble {\bf 33}, 3, 65-82 (1984).
\smallskip
{\hfill \break}
\Ch\ { R. Charles} and { Y. Diakite},
Sur les varietes d'algebres de { Lie} de dimension $\le 7$.
J. Algebra {\bf 91}, 53-63 (1984).
\smallskip
{\hfill \break}
\Co\ { C. W. Conatser},
Contractions of low-dimensional real { Lie} algebras.
J. Math. Phys. {\bf 13}, 196-203 (1972).
\smallskip
{\hfill \break}
\Est\ { F. B. Estabrook, H. D. Wahlquist} and { C. G. Behr},
Dyadic analysis of spatially homogeneous world models.
J. Math. Phys. {\bf 9}, 497-504 (1968).
\smallskip
{\hfill \break}
\Gri\ { D. R. Grigore} and { O. T. Popp},
On the classification of { Lie} subalgebras and { Lie} subgroups.
J. Math. Phys. {\bf 32}, 33-39 (1991).
\smallskip
{\hfill \break}
\Gru\ { F. Grunewald} and { J. O'Halloran},
Varieties of nilpotent lie algebras of dimension less than six.
J. Algebra {\bf 112}, 315-325 (1988).
\smallskip
{\hfill \break}
\Hud\ { P. L. Huddleston},
In\"on\"u-Wigner contractions of the real four-dimensional { Lie} algebras.
J. Math. Phys. {\bf 19}, 1645-1649 (1978).
\smallskip
{\hfill \break}
\In\ { E. In\"on\"u} and { E. P. Wigner},
On the contraction of groups and their representations.
Proc. Nat. Acad. Sci. USA {\bf 39}, 510-524 (1953).
\smallskip
{\hfill \break}
-, On a particular type of convergence to a singular matrix.
Proc. Nat. Acad. Sci. USA {\bf 40}, 119-121 (1954).
\smallskip
{\hfill \break}
\Ki\ { A. A. Kirillov} and { Yu. A. Neretin},
The variety $A_n$ of $n$-dimensional { Lie} algebra structures.
Transl., Ser. (2), Amer. Math. Soc. {\bf 137}, 21-30 (1987);
translation from: Some questions in modern analysis, Work. Collect.,
Moskva 1984, 42-56 (1984).
\smallskip
{\hfill \break}
{\newpage } \noindent
\Kr\ { D. Kramer, H. Stephani, M. Mac Callum} and { E. Herlt},
Exact Solutions of Einstein's Field Equations.
Wissenschaften, Berlin 1980.
\smallskip
{\hfill \break}
\Lan\ { L. D. Landau} and { E. M. Lifschitz},
Lehrb. d. theor. Phys., Bd. II,
12. Aufl., pp. 457 ff.
Akademie-Verlag, Berlin 1992.
\smallskip
{\hfill \break}
\Lee\ { H. C. Lee},
Sur les groupes de { Lie} r\'eels \`a trois param\`etres.
J. Math. Pures et Appl. {\bf 26}, 251-267 (1947).
\smallskip
{\hfill \break}
\Lie\ { S. Lie}, Differentialgleichungen, Kap. 21. Chelsea,
Leipzig 1891.
\smallskip
{\hfill \break}
\bLie\ -,
Theorie der Transformationsgruppen, Bd. III, Abtheilung VI,
Kap. 28: Allgemeines \"uber die Zusammensetzung $r$-gliedriger
Gruppen, \S\S 136-137, p. 713-732.
Teubner, Leipzig 1893.
\smallskip
{\hfill \break}
\Mac\ { M. A. H. Mac Callum},
A class of homogeneous cosmological models III: Asymptotic behaviour.
Commun. Math. Phys. {\bf 20}, 57-84 (1971).
\smallskip
{\hfill \break}
\Mag\ { L. M. Magnin},
Sur les algebres de { Lie} nilpotent de dim $\leq 7$.
J. G. P. {\bf 3}, 1, 119-134 (1986).
\smallskip
{\hfill \break}
\Mo\ { V. V. Morozov},
Classification of nilpotent { Lie} algebras of 6$^{th}$ order.
Izv. Vyssh. Uchebn. Zaved. Mat. {\bf 4}, 161-171 (1958).
\smallskip
{\hfill \break}
\aMu\ { G. M. Mubarakzyanov}, On solvable { Lie} algebras. (Russian)
Izv. Vys\v s. U\v cebn. Zavedeni\u\i\ Mat. 1963, no. 1 (32), 114-123.
\smallskip
{\hfill \break}
\bMu\ -, Classification of real structures of { Lie}
algebras of fifth order. (Russian)
Izv. Vys\v s. U\v cebn. Zavedeni\u\i\ Mat. 1963, no. 3 (34), 99-106.
\smallskip
{\hfill \break}
\cMu\ -, Classification of solvable { Lie} algebras
of sixth order with a non-nilpotent basis element. (Russian)
Izv. Vys\v s. U\v cebn. Zavedeni\u\i\ Mat. 1963, no. 4 (35), 104-116 (1963).
\smallskip
{\hfill \break}
\Ne\ { Yu. A. Neretin},
Estimate of the number of parameters assigning an $n$-dimensional algebra.
(Russian)
Isv. Acad. Nauk SSSR, Ser. Mat. {\bf 51}, 306-318 (1987).
\smallskip
{\hfill \break}
\PaSWZ\ { J. Patera, R. T. Sharp, P. Winternitz} and { H. Zassenhaus},
Invariants of real low dimension { Lie} algebras.
J. Math. Phys. {\bf 17}, 986-994 (1976).
\smallskip
{\hfill \break}
\PaW\ { J. Patera} and { P. Winternitz},
Subalgebras of real three- and four-dimensional { Lie} algebras.
J. Math. Phys. {\bf 18}, 1449-1455 (1977).
\smallskip
{\hfill \break}
\Pe\ { A. S. Petrov}, { Einstein}r\"aume.
Akademie Verlag, Berlin 1964.
\smallskip
{\hfill \break}
\aRa\ { M. Rainer}, Topology of the space of 4-dimensional real
{ Lie} Algebras. Preprint 93/2, Math. Inst. Univ. Potsdam (1993).
\smallskip
{\hfill \break}
\bRa\ { M. Rainer}, Orientation Duality of 4-dimensional Real
{ Lie} Algebras.
Preprint 93/8, Math. Inst. Univ. Potsdam (1993).
\smallskip
{\hfill \break}
\Ri\ { W. Rinow}, Topologie.
Wissenschaften, Berlin 1975.
\smallskip
{\hfill \break}
\Ro\ { M. Romdhani},
Classification of real and complex nilpotent { Lie} algebras of dimension 7.
Linear Multilin. Algebra {\bf 24}, 167-189 (1989).
\smallskip
{\hfill \break}
\Sal\ { E. I. Saletan},
Contraction of { Lie} groups.
J. Math. Phys. {\bf 2}, 1-21 (1961).
\smallskip
{\hfill \break}
\San\ { R. M. Santilli}, Lie-isotopic lifting of Lie symmetries,
I: General considerations. Hadronic J. {\bf 8} 25-35 (1985).
\smallskip
{\hfill \break}
-, Lie-isotopic lifting of Lie symmetries,
II: Lifting of Rotations. Hadronic J. {\bf 8} 36-51 (1985).
\smallskip
{\hfill \break}
\aSch\ { H.-J. Schmidt},
Inhomogeneous cosmological models containing homogeneous inner hypersurface
geometry. Changes of the { Bianchi} type.
Astron. Nachr. {\bf 303}, 227-230 (1982).
\smallskip
{\hfill \break}
\bSch\ -,
On Geroch's limit of space-times and its relation to a new topology in the
space of { Lie} groups.
J. Math. Phys. {\bf 28}, 1928-1936 (1987);
and Addendum in J. Math. Phys. {\bf 29}, 1264 (1988).
\smallskip
{\hfill \break}
\Se\ { I. E. Segal},
A class of operator algebras which are determined by groups.
Duke Math. J. {\bf 18}, 221-265 (1951).
\smallskip
{\hfill \break}
\Smrz\ { P. K. Smrz},
Relativity and deformed { Lie} groups.
J. Math. Phys. {\bf 19}, 2085-2088 (1978).
\smallskip
{\hfill \break}
\aTur\ { P. Turkowski}, Low-dimensional { Lie} algebras.
J. Math. Phys. {\bf 29}, 2139-2144 (1988).
\smallskip
{\hfill \break}
\bTur\ -,
Solvable { Lie} algebras of dimension six.
J. Math. Phys. {\bf 31}, 1344-1350 (1990).
\smallskip
{\hfill \break}
\cTur\ -,
Structure of real { Lie} algebras.
Linear Algebra Appl. {\bf 171}, 197-212 (1992).
\smallskip
{\hfill \break}
\Um\ { K. A. Umlauf}, \"uber die Zusammensetzung der endlichen
continuierlichen Transformationsgruppen, insbesondere der Gruppen
vom Range Null.
Breitkopf u. H\"artel, Leipzig 1891.
\smallskip
{\hfill \break}
\Vra\ { G. Vranceanu}, Lecons de geom\'etrie diff\'erentielle,
p. 105-111. Bucarest 1947.
\smallskip
{\hfill \break}
\end{document}
|
\section*{TWISTOR PHASE SPACE DYNAMICS \break AND THE LORENTZ FORCE EQUATION.}
\vskip 10 pt
\leftline{By}
\leftline{Andreas Bette,}
\leftline{Stockholm University,}
\leftline{Department of Physics,}
\leftline{Box 6730,}
\leftline{S-113 85 STOCKHOLM,}
\leftline{SWEDEN.}
\vskip 10 pt
\leftline{fax +46-8347817 att. Andreas Bette.}
\leftline{e-mail: $<$ab@vanosf.physto.se$>$.}
\vskip 50 pt
\centerline{ABSTRACT}
\vskip 15 pt
Using Lorentz force equation as an input a Hamiltonian mechanics on
the non-projective two twistor phase space TxT is formulated.
\vskip 10pt
Such a construction automatically reproduces dynamics of the intrinsic
classical relativistic spin.
\vskip 10pt
The charge appears as a dynamical variable.
\vskip 10pt
It is also shown that if the classical relativistic spin function on
TxT vanishes, the natural conformally invariant symplectic structure
on TxT reduces to the natural symplectic structure on the cotangent
bundle of the Ka{\l}u{\.z}a-Klein space.
\vfill
\eject
\section{INTRODUCTION.}
The classical motion of a relativistic electrically charged massive
and spinning particle exposed to an external
electromagnetic
field is, in Minkowski space, described by the Lorentz-Dirac (LD)
force equation and by the so called Bargmann, Michel, Telegdi (BMT)
equation for the intrinsic angular momentum (the spin).
If we by $X^{a}$, $P_{a}$, $S_{a}$, $F_{ab}$, $m^{2}:=P^{b}P_{b}$, $e$
and $g$ denote the four-position, the four-momentum, the Pauli-Luba{\'n}ski
four-vector, the external electromagnetic field tensor, the mass squared,
the charge and the gyromagnetic ratio of the particle then
these Poincar{\'e} covariant equations may be written as follows:
$$\dot X^{a} = P^{a},\eqno (1.1)$$
$$\dot P_{a} = eF_{ab}P^{b} + D_{a},\eqno (1.2)$$
$$\dot S_{a} = {ge\over 2}F_{ab}S^{b}+{ge\over 2m^{2}}(F_{ik}S^{i}P^{k})P_{a}
-{1 \over m^{2}}({\dot P_{k}}S^{k})P_{a}\eqno (1.3)$$
\vskip 10pt
where
$$P_{a}S^{a}=0 \eqno (1.4)$$
and
$$D_{a}P^{a}=0 \eqno (1.5).$$
\vskip 10pt
$D_{a}$ is a small space-like
correction four-vector (small compared with the space-like four-vector
$eF_{ab}P^{b}$) containing higher derivatives
of the external electromagnetic field $F_{kl}$, $F^{*}_{kl}$
and terms
nonlinear
in the spin variable $S^{i}$ [1,2].
\vskip 15pt
When the particle
forms (a classical limit of) an electron and the
radiation damping effects are neglected the value of $g$ equals $2$
(the Dirac value).
\vskip 15pt
The equations (1.1) - (1.5) are such that
the mass squared and the
spin squared of the particle:
$$m^{2}:=P_{a}P^{a} \ \ \ \ \ \ s^{2}:=-{1 \over m^{2}}S_{a}S^{a}\eqno (1.6)$$
\vskip 10pt
are constants of the motion.
\vskip 15pt
The dot in (1.1) - (1.3) denotes differentiation with respect
to a real parameter $l$
which is, by virtue of (1.1), linearly related to the proper
time ${\tau}$ of the particle by:
$${\tau}= {\pm}ml+{\tau_{0}}\eqno (1.7).$$
\vskip 10pt
${\tau_{0}}$ is an arbitrary real number representing the freedom
of choice of the origin of the proper time.
\vskip 15pt
$F_{ab}$ denotes the value of the
external electromagnetic field evaluated at the particle's four-position
$X^{a}$.
Consequently, the four-position coincides with
the location of the
charge $e$.
\vskip 15pt
In this paper we assume that $D_{a}=0$ in (1.2) and then
examine (1.1) and (1.2)
using two distinct twistors as variables.
\vskip 15pt
This analysis
will automatically
produce the BMT equation in (1.3) with $g=2$ [4].
\vskip 15pt
In the next section we give a physical interpretation to
the sixteen variables corresponding to a point in the space
of two twistors TxT [3,4].
\vskip 15pt
In section three the free particle
symplectic potential on TxT is expressed using these physical variables. The
non-uniqeness
of choice of the free particle Hamiltonian is disscused.
\vskip 10pt
The free particle equations of motion given as a canonical flow in the
phase space of two twistors are presented in twistors' Weyl spinor coordinates
as well as in Poincar{\'e} covariant physically interpretable coordinates.
These has been presented before [3] in a somewhat preliminary shape.
\vskip 15pt
In section four a deformed Poincar{\'e}
covariant symplectic structure and a deformed Poincar{\'e} scalar
Hamiltonian function on TxT are presented.
The new Poincar{\'e} covariant
flow in TxT
canonical with respect to the deformed symplectic structure
and generated by the deformed Hamiltonian
reproduces (1.1) - (1.4) (with $D_{a}=0$ and $g=2$)
and also produces certain additional equations of motion.
The latter arise because TxT is sixteen dimensional while the number
of independent variables describing the particle
according to (1.1) - (1.4) is only twelve (the four-position,
the four-momentum, the Pauli-Luba{\'n}ski spin four-vector
fulfilling (1.4) and the charge).
Our attempt to interpret physically the remaining
four variables is presented already in section two.
However, a (partial) confirmation of the correctness of
these tentative identifications
is provided first
when the interaction with an external electromagnetic
field
is "switched" on. This is done in section four.
\vskip 15pt
A first version of the material contained in section four
appeared in [5] where the electric charge was not defined as a
dynamical variable. This weakness of the model is removed in section
four of the present paper.
\vskip 15pt
In the appendix the formal proof of the statements made in section four
is presented.
\vskip 15pt
Upper case latin letters with lower case greek indices denote twistors.
Upper case latin letters with lower case latin indices denote four-vectors
and four-tensors.
Lower case greek letters with upper case latin indices (either primed
or unprimed) denote Weyl spinors.
The Minkowski metric has the signature $+---$.
\section {PHYSICAL VARIABLES IDENTIFIED \break AS FUNCTIONS ON T$\Delta$T.}
The symbol TxT usually denotes the direct product of two twistor spaces.
However, in our investigations, we will not be using the whole of TxT
but rather T$\Delta$T which from now on will denote the space TxT
less its diagonal i.e.
T$\Delta$T := $\{$TxT - $\{$(t, t) $\epsilon$ TxT; t $\epsilon$ T$\} \}$.
\vskip 15pt
The twistor coordinates of a point in T$\Delta$T will be expressed in
terms of two Weyl spinors:
$$Z^{\alpha} = (\omega^{A},\ \pi_{A^\prime})
\qquad and \qquad W^{\alpha} = (\lambda^{A},\ \eta_{A^\prime})\eqno (2.1)$$
\vskip 10pt
or dually (complex conjugation):
$$\overline Z_{\alpha} = ( \overline \pi_{A},\ \overline \omega^{A^\prime})
\qquad and \qquad \overline W_{\alpha} =
(\overline \eta_{A},\ \overline \lambda^{A^\prime})\eqno(2.2).$$
\vskip 10pt
Using these two twistors and their twistor conjugates four
independent conformally (SU(2,2))
scalar functions may be formed on T$\Delta$T [4,7,8]:
$$s_{1} = Z^{\alpha}\overline Z_{\alpha} \qquad and \qquad
s_{2} = W^{\alpha}\overline W_{\alpha}\eqno (2.3)$$
$$a = Z^{\alpha}\overline W_{\alpha} \qquad and \qquad
\overline a = W^{\alpha}\overline Z_{\alpha}\eqno (2.4).$$
\vskip 10pt
In addition, the following two Poincar\'e scalar functions
may also be defined on T$\Delta$T:
$$f = \pi^{A^\prime}\eta_{A^\prime} \qquad and \qquad
{\overline f} = \overline \pi^{A}\overline \eta_{A}\eqno (2.5).$$
\vskip 10pt
The scalar functions introduced above may be represented by six real
valued functions on T$\Delta$T given by:
$$e = s_{1} + s_{2} \qquad and \qquad k = s_{1} - s_{2} \eqno (2.6)$$
$$\mid a\mid \qquad and \qquad \vartheta = arga = -arg{\overline a}\eqno
(2.7)$$
$$\mid f\mid \qquad and \qquad \varphi = argf = -arg{\overline f}\eqno (2.8).$$
\vskip 15pt
Below, Poincar{\'e} covariant
functions
on T$\Delta$T will be identified as
physical quantities
according to the following recipe [4,5]
(we employ here the abstract index notation according to Penrose [6]):
$$P_{a} := \pi_{A^\prime}{\overline\pi}_A +
\eta_{A^\prime}{\overline\eta}_A\eqno (2.9)$$
\vskip 10pt
will denote a massive particle's four-momentum
expressed as a sum of
the four-momenta of its two massless parts.
\vskip 15pt
$$X^{a} := {1\over 2}(Z^{a} + {\overline Z}^{a})\qquad where \qquad
Z^{a} := {i\over f}(\omega^{A}\eta^{A^\prime}-\lambda^{A}\pi^{A^\prime})
\eqno (2.10)$$
\vskip 10pt
will denote a massive particle's four-position in Minkowski space time.
\vskip 15pt
A massive particle's Pauli-Luba{\'n}ski spin four-vector will be given
by:
$$S_{a} := {k\over 2}(\pi_{A^\prime}{\overline\pi}_A - \eta_{A^\prime}
{\overline\eta}_A) + a\eta_{A^\prime}{\overline\pi}_A +
{\overline a}\pi_{A^\prime}{\overline\eta}_A\eqno (2.11).$$
\vskip 10pt
The definition in (2.11)
is dictated by the assumption that
a massive particle's four angular
momentum should be a sum of the four angular
momenta of its two massless parts (see e.g. [3]).
\vskip 15pt
Note that $P_{a}$ and $S_{a}$ are automatically orthogonal to each other
i.e. we always have $P_{a}S^{a}=0$.
\vskip 15 pt
{}From the above it follows that the imaginary part of $Z^{a}$:
$$Y^{a} = {1\over 2i}(Z^{a} - {\overline Z}^{a})=\eqno (2.12)$$
\vskip 10pt
may be written as:
$$= {1\over 2f\overline f}\Big[(a\eta^{A^\prime}{\overline\pi}^A +
{\overline a}\pi^{A^\prime}{\overline\eta}^A) -
s_{1}\eta^{A^\prime}{\overline\eta}^A -
s_{2}\pi^{A^\prime}{\overline\pi}^A\Big] =\eqno (2.13)$$
\vskip 10pt
or as:
$$={1\over 2f\overline f}(S^{a} - {e\over 2} P^{a})\eqno (2.14).$$
\vskip 10pt
{}From the definitions above it also follows that on T$\Delta$T the mass
function of the particle is given by
$$m={\sqrt 2} \mid f\mid \eqno (2.15)$$
\vskip 10pt
while its spinfunction by:
$$s=\sqrt {{1\over 4} k^{2} + {\mid a\mid}^{2}}\eqno (2.16).$$
\vskip 10pt
A space-like plane spanned by two mutually orthogonal unit four-vector
valued functions on T$\Delta$T orthogonal to $S_{a}$ and $P_{a}$:
$$E_{a} := {i \over (m {\mid a \mid})}
(a\eta_{A^\prime}{\overline\pi}_A -
{\overline a}
\pi_{A^\prime}{\overline\eta}_A)
\eqno (2.17).$$
$$F_{a} := {1 \over (sm {\mid a \mid})}
[{k \over 2}(a\eta_{A^\prime}{\overline\pi}_A +
{\bar a}
\pi_{A^\prime}{\overline\eta}_A) -
{\overline a}a
(\pi_{A^\prime}{\overline\pi}_A - \eta_{A^\prime}
{\overline\eta}_A)]
\eqno (2.18)$$
\vskip 10pt
may be thought of as a polarization plane rigidly
attached to the massive particle at its four-position
$X^{a}$ in Minkowski space [3,4].
\vskip 15pt
In effect, all four four-vectors
${P_{a}/m}$, ${S_{a}/(sm)}$,
$E_{a}$ and $F_{a}$ span an orthonormal tetrad rigidly attached
to the particle at its four-position
$X^{a}$ in Minkowski space. The number of variables represented by the
functions defining this tetrad
is six, the number of variables represented by
the scalar functions is also six, while the four-position represents
four variables; sixteen variables altogether.
\vskip 15pt
With these identifications the
inverse relations expressing twistor coordinates in (2.1) - (2.2)
as functions of the introduced Poincar{\'e} covariant physical
variables and the scalars in (2.6) - (2.8) (note that according
to (2.15) and (2.16) two of these scalars have a clear physical
interpretation) are almost immediate.
\vskip 15pt
The spinor $\pi_{A^\prime}$ up to its phase is given by:
$$\pi_{A^\prime}{\overline\pi}_A ={{1 \over 2}(P_{a} + {k \over 2s^{2}}S_{a} -
{m {\mid a \mid} \over s}F_{a})} \eqno (2.19),$$
\vskip 10pt
the spinor $\eta_{A^\prime}$ up to its phase is given by:
$$\eta_{A^\prime}{\overline\eta}_A = {{1 \over 2}(P_{a} -
{k \over 2s^{2}}S_{a} +
{m {\mid a \mid} \over s}F_{a})} \eqno (2.20),$$
\vskip 10pt
while the phase $\alpha$ of the spinor $\pi_{A^\prime}$ is given by:
$$\alpha={1 \over 2}(argf + arg a) = {1 \over 2}(\varphi + \vartheta)
\eqno (2.21),$$
\vskip 10pt
and the phase $\beta$ of the spinor $\eta_{A^\prime}$ by:
$$\beta = {1 \over 2}(argf - arg a) = {1 \over 2}(\varphi - \vartheta)
\eqno (2.22).$$
\vskip 10pt
The relations in (2.21) and (2.22) follow from (2.5) and from
the fact that the conformal
complex valued scalar $a$ in (2.4) may be written as:
$$a=-2Y^{A{A^\prime}}{\overline \eta}_{A}{\pi_{A^\prime}} \eqno (2.23)$$
\vskip 10pt
where $Y^{a}$ is a real four-vector valued function on T$\Delta$T introduced in
(2.12).
\vskip 15pt
The remaining spinors are given by (see (2.14) and (2.15)):
$$\omega^{A} = iX^{AA^{\prime}}\pi_{A^\prime} -
{1 \over m^{2}}(S^{AA^{\prime}}\pi_{A^\prime} -
{e \over 2} P^{AA^{\prime}}\pi_{A^\prime})
\eqno (2.24)$$
and
$$\lambda^{A} = iX^{AA^{\prime}}\eta_{A^\prime}
-{1 \over m^{2}}(S^{AA^{\prime}}\eta_{A^\prime} -
{e \over 2} P^{AA^{\prime}}\eta_{A^\prime}) \eqno (2.25).$$
\section {THE FREE PARTICLE MOTION.}
The two twistor space T$\Delta$T possesses a natural
(free particle) symplectic structure given by [7,8]:
$$\Omega_{0} = i(dZ^{\alpha}\wedge d\overline Z_{\alpha} + dW^{\alpha}\wedge
d\overline W_{\alpha}) \eqno (3.1).$$
\vskip 10pt
$\Omega_{0}$ may be regarded as exterior derivative of a
one-form $\gamma_{0} \ (\Omega_{0} = d\gamma_{0})$ given by:
$$\gamma_{0} = {i\over 2}(Z^{\alpha}d\overline Z_{\alpha} -
\overline Z_{\alpha}dZ^{\alpha} + W^{\alpha}d\overline W_{\alpha} -
\overline W_{\alpha}dW^{\alpha})\eqno (3.2).$$
\vskip 10pt
Using the introduced Poincar{\'e} covariant physical functions on T$\Delta$T,
$\gamma_{0}$ may also
be written as:
$$\gamma_{0} = P_{j}dX^{j} + {1\over 2}ed\varphi - {1\over 2}kd\vartheta +
({k^{2} \over 4s}F_{j} +
{{\mid a \mid}k \over 2ms^{2}}S_{j} +
{{\mid a \mid} \over m}P_{j})dE^{j}
\eqno (3.3)$$
\vskip 10pt
or equivalently
$$\gamma_{0} = P_{j}dX^{j} + {1\over 2}ed\varphi - {1\over
2}kd\vartheta + {k \over 2m}(iM_{j}d{\bar M}^{j} - i{\bar
M}_{j}dM^{j}) + $$
$$+{i{\bar a} \over m^{2}}M_{j}dP^{j}- {ia \over
m^{2}}{\bar M}_{j}dP^{j} \eqno (3.4)$$
\vskip 10pt
where $M_{j}$ is a complex null four-vector valued function on T$\Delta$T
given by:
$$M_{a}:=\pi_{A^{\prime}}{\bar \eta}_{A} \eqno (3.4a).$$
\vskip 10pt
{}From (3.3) or (3.4) we notice a remarkable fact that for $a=k=0$
i.e. for the vanishing value of the spin function on T$\Delta$T, the
conformally invariant symplectic potential $\gamma_{0}$ in (3.2)
(and thereby also the symplectic structure $\Omega_{0}$ in (3.1))
reduces to the natural symplectic potential
(while $\Omega_{0}$ reduces to the natural symplectic structure)
on the cotangent bundle of the Ka{\l}u{\.z}a-Klein space.
This suggests that $e$ should be identified with the electric
charge of the particle.
\vskip 15pt
To generate the free motion of a massive particle built up of
the two twistors we used in [3] a Hamiltonian:
$$H_{0_1} = m^{2} + s^{2}\eqno (3.5)$$
\vskip 10pt
and a somewhat modified one in [4]:
$$H_{0_2} = {1\over 2}(m^{2} + s^{2}) \eqno (3.6).$$
\vskip 10pt
Any such a change is of no importance as long as
$H_{0}$ is a function on T$\Delta$T such that:
$$H_{0} = H_{0}(m,\ s) \eqno (3.7).$$
\vskip 10pt
The flow will always correspond to a free particle motion in Minkowski space.
In fact
any function such that:
$$H_{0} = H_{0}(m,\ s,\ k,\ e)\eqno (3.8)$$
\vskip 10 pt
describes a free particle. As $m$, $s$, $k$ and $e$ are mutually
(Poisson) commuting functions the different choices of $H_{0}$
may correspond to different motions of the internal physical variables
represented by
$\varphi$ and
$\vartheta$.
\vskip 10pt
But in most cases different choices of $H_{0}$ simply correspond to a
reparametri- zation of the canonical flow lines.
\vskip 15pt
At this non-quantum level there is thus quite a large freedom
of choice of the free particle
Hamiltonian $H_{0}$. On the quantum level of this approach
one should, on the other hand, expect essential differences
depending on the choice of ${\hat H}_{0}$.
\vskip 10pt
In this paper, for simplicity, we choose $H_{0}$ as:
$$H_{0}:={1 \over 2}m^{2} + (s^{2} - {1\over 4}e^{2}) \eqno (3.9)$$
\vskip 10pt
which written out in terms of the introduced scalar functions yields:
$$H_{0} := f{\overline f} + {1\over 4}k^{2} + a{\overline a} - {1\over 4}e^{2}
= f{\overline f} - s_{1}s_{2} + a{\overline a} \eqno (3.10).$$
\vskip 10pt
The chosen $H_{0}$ and $\Omega_{0}$ in (3.1) generate the following equations
of motion in T$\Delta$T:
$$\dot\omega^A = -if\overline\eta^A + ia\lambda^A - is_{2}\omega^A
\eqno (3.11)$$
$$\dot\pi_{A^\prime} = ia\eta_{A^\prime} - is_{2}\pi_{A^\prime} \eqno (3.12)$$
$$\dot\lambda^A = if\overline\pi^A + i{\overline a}\omega^A -
is_{1}\lambda^A \eqno (3.13)$$
$$\dot\eta_{A^\prime} = i{\overline a}\pi_{A^\prime} - is_{1}\eta_{A^\prime}
\eqno (3.14)$$
\vskip 10pt
and their complex conjugates (c.c.).
\vskip 15pt
The above equations, written out using functions representing
the physical variables
as previously identified, read:
$${\dot e}= 0 \qquad and \qquad {\dot k} = 0 \eqno (3.15)$$
$${\dot \varphi} = - e \qquad and \qquad {\dot \vartheta} = 0 \eqno (3.16)$$
$${\dot X}^{a} = P^{a} \eqno (3.17)$$
$${\dot P}_{a} = 0 \qquad and \qquad {\dot S}_{a} = 0 \eqno (3.18)$$
$${\dot E}_{a} = 2s F_{a} \eqno (3.19)$$
$${\dot F}_{a} = -2s E_{a} \eqno (3.20).$$
\vskip 10pt
{}From (3.19), (3.20) and (1.7) it follows that, with our choice of $H_{0}$,
the introduced polarization plane rigidly
attached to the particle rotates with an angular velocity equal
to $(2s/m)$ [3].
\section {MOTION IN AN EXTERNAL ELECTRO \break MAGNETIC FIELD.}
In this section we identify the function $e$ on T$\Delta$T with
the electric charge of the particle. The deformed Poincar{\'e}
covariant symplectic potential on T$\Delta$T we define as:
$$\gamma = \gamma_{0}+eA_{j}dX^{j} \eqno (4.1)$$
\vskip 10pt
where $X^{a}$ is a four vector-valued function on T$\Delta$T given by (2.10)
and
where $A_{j}=A_{j}(X^{a})$
denotes an external electromagnetic four-potential. $A_{j}$ is in this way
a four-vector valued function defined on T$\Delta$T.
$\gamma_{0}$ is given by (3.3) (or equivalently
by (3.2) or (3.4)).
\vskip 15pt
The external derivative of $\gamma$ gives us the deformed symplectic
structure on T$\Delta$T:
$$\Omega = \Omega_{0} + de\wedge dX^{j}A_{j}
+ {1\over 2}eF_{jk}dX^{j}\wedge dX^{k} \eqno (4.2)$$
\vskip 10pt
where $F_{jk}=F_{jk}(X^{a})$ denotes the
electromagnetic field tensor formed from
$A_{j}$. $\Omega_{0}=d\gamma_{0}$.
\vskip 10pt
Note that for $a=k=0$, $\gamma$ and thereby $\Omega$ may be regarded
as a deformation of the
natural symplectic potential and natural symplectic structure on
the cotangent bundle of the Ka{\l}u{\.z}a-Klein space.
\vskip 15pt
As the deformed Hamiltonian function on T$\Delta$T we take:
$$H = H_{0} + {e \over m^{2}}F^{*}_{jk}S^{j}P^{k} \eqno (4.3)$$
\vskip 10pt
where $F^{*}_{jk}=F^{*}_{jk}(X^{a})$
represents on T$\Delta$T the dual of the external
electromagnetic tensor field.
\vskip 15pt
It is shown in the appendix that, with respect to $\Omega$, $H$
generates a Poincar{\'e} covariant canonical flow
in T$\Delta$T provided Maxwell's empty space
equations are fulfilled at the location of the particle:
$$F^{*}_{\lbrack jk,n\rbrack} = 0 \eqno (4.4).$$
\vskip 10pt
For future reference we note that
using (2.12), (2.14), (2.15) and the skew symmetry of the dual of the
external electro-magnetic field tensor the generating function $H$ may
also be written as:
$$H = H_{0} + eF^{*}_{ik}Y^{i}P^{k} \eqno (4.5).$$
\vskip 10pt
Expressed in twistor coordinates the flow canonical with respect to
$\Omega$ and generated by $H$ is given by the following equations of
motion (see proof in the appendix):
$$\dot\omega^A = -if\overline\eta^A + ia\lambda^A - is_{2}\omega^A+$$
$$ + e\mu^A_{\ B}Y^{B{B^\prime}}\pi_{B^\prime} +
ieX^{AA^\prime}{{\overline \mu}_{A^\prime}\ ^{B^{\prime}}}\pi_{B^\prime}
+ iC\omega^A \eqno (4.6)$$
$$\dot\pi_{A^\prime} = ia\eta_{A^\prime} - is_{2}\pi_{A^\prime} +
e{{\overline \mu}_{A^\prime}\ ^{B^{\prime}}}\pi_{B^\prime} +
iC\pi_{A^\prime} \eqno (4.7)$$
$$\dot\lambda^A = if\overline\pi^A + i{\overline a}\omega^A - is_{1}\lambda^A
+ $$
$$+e\mu^A_{\ B}Y^{B{B^\prime}}\eta_{B^\prime} +
ieX^{AA^\prime}{{\overline \mu}_{A^\prime}\ ^{B^{\prime}}}\eta_{B^\prime} +
iC\lambda^A \eqno (4.8)$$
$$\dot\eta_{A^\prime} = i{\overline a}\pi_{A^\prime} - is_{1}\eta_{A^\prime} +
e{{\overline \mu}_{A^\prime}\ ^{B^{\prime}}}\eta_{B^\prime} + iC\eta_{A^\prime}
\eqno (4.9)$$
where
$$C = (F^{*}_{ik}Y^{i}P^{k}-A^{i}P_{i}) \eqno (4.10)$$
\vskip 10pt
and where $\mu_{AB} = \mu_{AB}(X^{c})$ is a spinor field corresponding
to $F_{ab}=F_{ab}(X^{c})$ [6]:
$$\mu_{AB}={1 \over 2}F_{AA^{\prime}B}\ ^{A^{\prime}} \eqno (4.11).$$
\vskip 10pt
Conversely one has [6]:
$$F_{ab} = \mu_{AB}\epsilon_{{A^\prime}{B^\prime}} +
{\overline \mu}_{{A^\prime}{B^\prime}}\epsilon_{AB} \eqno (4.12)$$
$$F^{*}_{ab} = i{\overline \mu}_{{A^\prime}{B^\prime}}\epsilon_{AB} -
i\mu_{AB}\epsilon_{{A^\prime}{B^\prime}} \eqno (4.13).$$
\vskip 10pt
Written out in terms of the introduced Poincar{\'e} covariant physical
variables the above equations of motion read:
$$\dot X^{j} = P^{j} \qquad and \qquad \dot P_{j} = eF_{jk}P^{k} \eqno (4.14)$$
$$\dot S_{j} = eF_{jk}S^{k} \eqno (4.15)$$
$$\dot e=0 \qquad and \qquad \dot k=0 \eqno (4.16)$$
$$\dot \vartheta = 0 \eqno (4.17)$$
$$\dot \varphi = -e - 2P_{j}A^{j} + {2 \over m^{2}}F^{*}_{jk}S^{j}P^{k}
\eqno (4.18)$$
$${\dot E}_{j} = 2s F_{j} + e F_{kj}E^{k} \eqno (4.19)$$
$${\dot F}_{j} = -2s E_{j} + e F_{kj}F^{k} \eqno (4.20).$$
\vskip 15pt
As may be seen the equations in (4.14) and (4.15) are the same as
those in (1.1) - (1.3) (with $D_{j}=0$ and $g=2$) while the relation
in (1.4) is automatically fulfilled because of the way $S_{j}$ and
$P_{j}$ were defined in (2.9) and (2.11).
\vskip 15pt
The charge function $e$ appears as a dynamical variable and according to
(4.16) is a constant of motion.
\vskip 15pt
Conformally scalar functions $k$ and $\vartheta$
in (4.16) and (4.17) do not yet have any
clear physical interpretation. They form two ((Poisson) non-commuting)
constants of motion.
\vskip 15pt
The first two terms in (4.18) correspond to the Aharonov-Bohm effect
while the third term arises because of the non-vanishing intrinsic
spin of the particle.
\vskip 15pt
The motion of the polarization plane is given by (4.19) and (4.20).
\vskip 15 pt
Finally we note that the equations of motion in (4.6) - (4.9) may also
be written in a twistor covariant way i.e. entirely in terms of
$Z^{\alpha}$ and $W^{\alpha}$:
$${\dot Z}^{\alpha} = (if + l_{1}f)
I^{\alpha \beta}{\bar W}_{\beta}
+(ia - {\bar c}_{2})W^{\alpha} - (is_{2} + iC + {\bar c}_{3})Z^{\alpha}
- b I^{\alpha \beta}{\bar Z}_{\beta} \eqno (4.21)$$
$${\dot W}^{\alpha} = (if + l_{2}f)I^{\beta \alpha}{\bar Z}_{\beta}
+(i{\bar a} + {\bar c}_{1})Z^{\alpha} - (is_{1} + iC - {\bar c}_{3})W^{\alpha}
- {\bar b} I^{\beta \alpha}{\bar W}_{\beta} \eqno (4.22)$$
\vskip 10pt
where $I^{\alpha \beta}$ is the so called infinity twistor and where
$c_{1}$, $c_{2}$, $c_{3}$ are certain, conveniently chosen, complex
valued Poincar{\'e} scalar functions on T$\Delta$T describing the
external electromagnetic field ($e$ and ${\bar f}$ are defined in
(2.5) - (2.6)):
$$c_{1}={e{\mu^{AB}}{\bar \eta}_{A}{\bar \eta}_{B} \over {\bar f}}
\eqno (4.23)$$
$$c_{2}={e{\mu^{AB}}{\bar \pi}_{A}{\bar \pi}_{B} \over {\bar f}} \eqno (4.24)$$
$$c_{3}=-{e{\mu^{AB}}{\bar \pi}_{A}{\bar \eta}_{B} \over {\bar f}}
\eqno (4.25).$$
\vskip 15pt
In (4.21) - (4.22) $l_{1}$ and $l_{2}$ are real valued Poincar{\'e}
scalar functions on T$\Delta$T while b is a complex valued
Poincar{\'e} scalar function on T$\Delta$T given by (see (2.3) and
(2.4)):
$$l_{1}=-{1 \over m^{2}}[ac_{2} + {\bar a}{\bar c}_{2} + s_{1} (c_{3}
+ {\bar c}_{3})] \eqno (4.26),$$
$$l_{2}={1 \over m^{2}}[{\bar a}c_{1} + a {\bar c}_{1} + s_{2} (c_{3}
+ {\bar c}_{3})] \eqno (4.27),$$
$$b={1 \over m^{2}}[a({\bar c}_{3}-c_{3}) - ({\bar c}_{2}s_{2} +
c_{1}s_{1})] \eqno (4.28).$$
\vskip 15pt
The deformed symplectic potential in (4.1), the corresponding
symplectic structure in (4.2), the deformed Hamiltonian in (4.3) (or
(4.5)) may all be written in a twistor covariant way i.e. entirely
in terms of $Z^{\alpha}$ and $W^{\alpha}$ and the infinity twistor
$I^{\alpha \beta}$.
The arising expressions are, however, quite complicated and not especially
illuminating from the physical point of view. In this paper we
therefore omit their presentation.
\section {CONCLUSIONS AND REMARKS.}
In this paper we describe the dynamics of a relativistic charged
particle with spin in an external electro-magnetic field using
two-twistor phase space T$\Delta$T. We have shown that there exists
a Hamiltonian dynamics on T$\Delta$T which after passing
to space-time coordinates reproduces the Lorentz force dynamics
and Bargmann-Michel-Telegdi
dynamics (with $g=2$) and also indicates connection to the
Ka{\l}u{\.z}a-Klein space dynamics. Conversely, one can say that
there exists a sort of the square root of the Lorentz force dynamics
which is realized as a Hamiltonian dynamics on T$\Delta$T.
\vskip 15pt
It will be interesting to see how the quantized version of the above
formalism corresponds to the Dirac equation coupled to an external
electromagnetic field. It seems that the approach developed by A.
Odzijewicz and his group [9-12] would be of great value here.
\section {REFERENCES.}
\leftline{[1] F. Rohrlich, "Classical Charged Particles", Addison-Wesley, 1965,
sect. 7-4,}
\vskip 10pt
\leftline{[2] L.D. Landau and E.M. Lifshitz, "The Classical Theory of Fields",}
\leftline {Pergamon Press Ltd., 1985, sect. 76,}
\vskip 10pt
\leftline{[3] A. Bette, J. Math. Phys., Vol. 25, No. 8, 2456-2460, August
1984,}
\vskip 10pt
\leftline{[4] A. Bette, Rep. Math. Phys., Vol. 28, No. 1, 133-140, 1989,}
\vskip 10pt
\leftline{[5] A. Bette, J. Math. Phys., Vol. 33, No. 6, 2158-2163, June 1992,}
\vskip 10pt
\leftline{[6] R. Penrose and W. Rindler, "Spinors and Space-Time"-(Cambridge}
\leftline{monographs on mathematical physics) Vol. 1: "Two-spinor calculus and}
\leftline{relativistic fields 1. Spinor analysis", Cambridge University Press,
1984,}
\vskip 10pt
\leftline{[7] K.P. Tod, Massive Spinning Particles and Twistor Theory, Doctoral
}
\leftline{Dissertation, Mathematical Institute, University of Oxford, Oxford,
1975,}
\vskip 10pt
\leftline{[8] K.P. Tod, Rep. Math. Phys., Vol. 7, No. 3, 339-346, 1977,}
\vskip 10pt
\leftline{[9] A. Odzijewicz, Comm. Math. Phys., Vol. 107, 561-575,
1986,}
\vskip 10pt
\leftline{[10] A. Karpio, A.Krysze{\'n}, A. Odzijewicz, Rep. Math. Phys.,}
\leftline{Vol. 24, No. 1, 65-80, 1986,}
\vskip 10pt
\leftline{[11] A. Odzijewicz, Comm. Math. Phys., Vol. 114, 577-597,
1988,}
\vskip 10pt
\leftline{[12] A. Odzijewicz, Comm. Math. Phys., Vol. 150, 385-413, 1992.}
\vskip 25pt
\section {APPENDIX; A FORMAL PROOF OF \break THE MAIN RESULT OF SECT. 4}
In order to prove that the Hamiltonian in (4.5) and the symplectic structure
$\Omega$ in (4.2) generate equations (4.6) - (4.9) which, in turn, imply
(4.14) - (4.20) we have to prove that:
$$V{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega=-dH \eqno (A.1)$$
\vskip 10pt
where $H$ is that in (4.5) $\Omega$ is that in (4.2) and where
$$ V = V_{0}+ V_{1} \eqno (A.2)$$
\vskip 10pt
where the vector-field $V_{0}$ according to (3.12)-(3.15) is given by:
$$V_{0}=(-if\overline\eta^A + ia\lambda^A - is_{2}\omega^A)
{\partial \over {\partial \omega^A}}
+ (ia\eta_{A^\prime} - is_{2}\pi_{A^\prime}){\partial \over {\partial
\pi_{A^\prime}}}+ $$
$$+(if\overline\pi^A + i{\overline a}\omega^A -
is_{1}\lambda^A){\partial \over {\partial \lambda^A}} +
(i{\overline a}\pi_{A^\prime} - is_{1}\eta_{A^\prime})
{\partial \over {\partial \eta_{A^\prime}}} + c.c. \eqno (A.3),$$
\vskip 10pt
or using the introduced four-vector variables (see (3.16)-(3.21)):
$$V_{0}=P^{j}{\partial \over {\partial X^{j}}} -
e {\partial \over {\partial \varphi}}
-2sE^{j}{\partial \over {\partial F^{j}}}
+ 2sF^{j}{\partial \over {\partial E^{j}}} \eqno (A.3a).$$
\vskip 10pt
The vector field $V_{1}$ is according to (4.6)-(4.9) given by:
$$V_{1}= (e\mu^A_{\ B}Y^{B{B^\prime}}\pi_{B^\prime} +
ieX^{AA^\prime}\overline \mu_{A^\prime}\ ^{B^{\prime}}\pi_{B^\prime} +
iC\omega^A) {\partial \over {\partial \omega^A}}+ c.c.+$$
$$+(e\overline\mu_{A^\prime}\ ^{B^{\prime}}\pi_{B^\prime} +
iC\pi_{A^\prime}){\partial \over {\partial \pi_{A^\prime}}}+ c.c.+$$
$$+(e\mu^A_{\ B}Y^{B{B^\prime}}\eta_{B^\prime} + ieX^{AA^\prime}
\overline\mu_{A^\prime}\ ^{B^{\prime}}\eta_{B^\prime} + iC\lambda^A)
{\partial \over {\partial \lambda^A}} + c.c.+$$
$$+(e{\overline\mu}_{A^\prime}\ ^{B^{\prime}}\eta_{B^\prime} +
iC\eta_{A^\prime}){\partial \over {\partial \eta_{A^\prime}}}+ c.c.
\eqno (A.4).$$
or using the introduced four-vector variables (see (4.14)-(4.20)):
$$V_{1}=eF_{jk}P^{k}{\partial \over {\partial P^{j}}} +
2C {\partial \over {\partial \varphi}} +
F^{kj}F_{k}{\partial \over {\partial F^{j}}}+
F^{kj}E_{k}{\partial \over {\partial E^{j}}} \eqno (A.4a).$$
\vskip 10pt
To facilitate the caculations
the inner product on the left hand side of (A.1) may be split
into a sum of partial inner products:
$$V{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega=V_{0}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega_{0}
+V_{0}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega_{1}+V_{1}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}
\Omega_{0}+V_{1}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega_{1} \eqno (A.5)$$
\vskip 10pt
where
$$\Omega = \Omega_{0} + \Omega_{1} \eqno (A.6)$$
$$\Omega_{0} = i(dZ^{\alpha}\wedge d\overline Z_{\alpha} +
dW^{\alpha}\wedge d\overline W_{\alpha}) \eqno (A.7)$$
and
$$\Omega_{1} = de\wedge dX^{i}A_{i} + {1\over 2}eF_{ik}dX^{i}\wedge dX^{k}
\eqno (A.8).$$
\vskip 10pt
By assumption, which may be checked by direct calculations, one has:
$$V_{0}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega_{0}=
-d({1\over 2}m^{2} + s^{2} - {1 \over 4}e^{2})
\eqno (A.9).$$
\vskip 10pt
Using the fact (see (3.16)-(3.21)) that:
$$V_{0}=P^{i}{\partial \over {\partial X^{i}}}+ \eqno (A.10)$$
\vskip 10pt
\centerline{ + terms in directions linearly independent of
${\partial \over {\partial X^{i}}}$}
\vskip 15 pt
and that $V_{0}$ has no component along ${\partial \over {\partial e}}$
one obtains by direct calculations:
$$V_{0}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}
\Omega_{1}=-A_{i}P^{i}de + eF_{ik}P^{i}dX^{k}
\eqno (A.11).$$
\vskip 10pt
Further, tedious spinor algebra manipulations yield:
$$V_{1}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega_{0}
= - Cde + e F_{ik}P^{k}dX^{i}
- e{F^{*}_{ik}}d(Y^{i}P^{k})=
- Cde + e F_{ik}P^{k}dX^{i}$$
$$ - d(e{F^{*}_{ik}}Y^{i}P^{k}) + eY^{i}P^{k}d{F^{*}_{ik}} +
(Y^{i}P^{k}{F^{*}_{ik}})de
\eqno (A.12).$$
\vskip 15pt
Using the fact that according to the
equations of motion the vector components of $V_{1}$ in the direction of
${\partial \over {\partial X^{i}}}$ and in the direction of
${\partial \over {\partial e}}$
are equal to zero one gets automatically:
$$V_{1}{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega_{1}= 0 \eqno (A.13).$$
\vskip 10pt
Putting together $(A.9)$, $(A.11)$, $(A.12)$, $(A.13)$ and
inserting $C=(Y^{i}P^{k}{F^{*}_{ik}}) - A_{i}P^{i}$ yields:
$$V{_{_{\_\!\_\!\_\!\_\!}}{\!|}}\Omega=
-d({1\over 2}m^{2} + s^{2} - {1 \over 4}e^{2})
- d(e{F^{*}_{ik}}Y^{i}P^{k}) + eY^{i}P^{k}d{F^{*}_{ik}}=
\eqno (A.14)$$
$$=-d({1\over 2}m^{2} + s^{2} - {1 \over 4}e^{2})
- d(e{F^{*}_{ik}}Y^{i}P^{k})=-dH \eqno (A.15)$$
\vskip 10pt
provided the last term in (A.14)
vanishes for all choices of $P_{i}$ and $Y_{i}$. That will always happen
if
the empty space Maxwell's equations:
$$F^{*}_{[ik,n]}=0 \eqno (A.16)$$
\vskip 10pt
are fulfilled at the location
of the particle i.e. at its four-position in Minkowski space.
\vskip 10pt
This completes the proof of our assertion.
\vskip 10pt
Note that the first pair of Maxwell's equations is satisfied by virtue
of the fact that the external electromagnetic field is given by means of
a four-potential $A_{j}$ in the expression for the symplectic one-form
$\gamma$. This automatically ensures that the symplectic structure
$\Omega$ is a closed two-form on T$\Delta$T.
\end{document}
\end
|
\section{Introduction}
The excited states of the Calogero-Sutherland model \cite{rSu}
and its relativistic model
(the trigonometric limit of the Ruijsenaars model) \cite{rR}
are described by the Jack polynomials \cite{rSt}
and their $q$-analog (the Macdonald polynomials) \cite{rM}, respectively.
Since the Jack polynomials coincide with
certain correlation functions of $\cW_N$ algebra \cite{rMY,rAMOS},
it is natural to expect that the Macdonald polynomials
are also realized by those of a deformation of $\cW_N$ algebra.
In a previous paper \cite{rSKAO},
we derived a quantum Virasoro algebra whose singular vectors are
some special kinds of Macdonald polynomials.
On the other hand,
E.~Frenkel and N.~Reshetikhin succeeded in constructing
the Poisson $\cW_N$ algebra
and its quantum Miura transformation
in the analysis of the $U_q(\widehat{sl_N})$ algebra at the critical level
\cite{rFR}.
Like the classical case \cite{rFL},
these two works, $q$-Virasoro and $q$-Miura transformation,
are essential to find and study a quantum $\cW_N$ algebra.
In this article,
we present a {$q$-${\cal W}_N$} algebra
whose singular vectors realize the general Macdonald polynomials.
This paper is arranged as follows:
In section 2, we define a quantum deformation of $\cW_N$ algebras
and its quantum Miura transformation.
The screening currents and a vertex operator are derived
in section 3 and 4.
A relation with the Macdonald polynomials is obtained in section 5.
Section 6 is devoted to conclusion and discussion.
Finally we recapitulate the $q$-Virasoro algebra and
the integral formula for the Macdonald polynomials in appendices.
\section{Quantum deformation of $\cW_N$ algebra}
We start with defining
a new quantum deformation of the $\cW_N$ algebra
by quantum Miura transformation.
\subsection{Quantum Miura transformation}
\newcommand\frenkel{
We found this commutation relation by comparing
the Poisson bracket in Frenkel-Reshetikhin's work \cite{rFR}
and the commutator in ours \cite{rSKAO}.
The oscillator $a_n$ used in \cite{rSKAO} is given by
$a_n = -n h^1_n p^{-n/2}/(1-t^n)$ and
$a_{-n} = n h^1_{-n} p^{n/2}(1+p^n)/(1-t^{-n})$ for $n>0$.
}
First we define fundamental bosons ${\h in}$ and ${\Qh i}$
for $i=1,2,\cdots,N$ and $n\in\bZ$ such that\footnote{\frenkel}
\begin{eqnarray}
[{\h in},{\h jm}] &\!\!=\!\!& -{1\/n}(1-q^n)(1-t^{-n})
{1-p^{(\,\delta_{ij}N-1)n}\/1-p^{Nn}} p^{Nn\theta(i<j)}
\,\delta_{n+m,0},\cr
[{\h i0},{\Qh j}] &\!\!=\!\!& \,\delta_{ij} - {1\/N},\qquad\quad
\sum_{i=1}^N p^{in} {\h in} = 0,\qquad\quad
\sum_{i=1}^N {\Qh i} =0,
\end{eqnarray}
with $q$, $t\equiv q^\beta\in\bC$ and $p\equiv q/t$.
Here $\theta(P)\equiv 1$ or $0$
if the proposition $P$ is true or false, respectively.
This bosons correspond to the weights of the vector representation
$h_i$ whose inner-product is $(h_i\cdot h_j)=(\,\delta_{ij}N-1)/N$.
Let us define fundamental vertices $\La_i(z)$ and
{$q$-${\cal W}_N$} generators $W^i(z)$ for $i=1,2,\cdots,N$ as follows:
\begin{eqnarray}
\La_i(z) &\!\!\equiv\!\!& :\Exp{ \sum_{n\neq 0}{\h in} z^{-n} }:
q^{\rb{\h i0}} p^{{N+1\/2}-i},\cr
W^i(zp^{1-i\/2}) &\!\!\equiv\!\!& \sum_{1\leq j_1<\cdots<j_i\leq N}
:\La_{j_1}(z) \La_{j_2}(zp^{-1}) \cdots \La_{j_i}(zp^{1-i}):,
\end{eqnarray}
and $W^0(z)\equiv 1$.
Here $:*:$ stands for the usual bosonic normal ordering such that
the bosons ${\h in}$ with non-negative mode $n\geq 0$ are in the right.
Note that
\begin{equation}
W^N(zp^{1-N\/2}) =
\,:\!\La_1(z) \La_2(zp^{-1}) \cdots \La_N(zp^{1-N})\!:\, = 1.
\end{equation}
If we take the limit $t\rightarrow 1$ with $q$ fixed,
the above generators reduce to those of Ref.\ \cite{rFR}.
These generators are obtained by the following quantum Miura transformation:
\begin{equation}
:\!\(p^{D_z} - \La_1(z)\) \(p^{D_z} - \La_2(zp^{-1})\) \cdots
\(p^{D_z} - \La_N(zp^{1-N})\)\!:\,
= \sum_{i=0}^N (-1)^i W^i(zp^{1-i\/2}) p^{(N-i)D_z},
\label{e:qMiura}
\end{equation}
with $D_z \equiv z{\partial\/\partial z}$.
Remark that $p^{D_z}$ is the $p$-shift operator such that
$p^{D_z} f(z) = f(pz)$.
\subsection{Relations of {$q$-${\cal W}_N$} generators}
Next we give the algebra of the above {$q$-${\cal W}_N$} generators.
Let $W^i(z) = \sum_{n\in\bZ} W^i_n z^{-n}$.
Let us define a new normal ordering $\:*\:$
for the {$q$-${\cal W}_N$} generators as follows:
\begin{eqnarray}
&\!\!&\!\!\!\!\!\!
\: W^i(rw)W^j(w)\: \cr
&\!\!\equiv\!\!&
\oint{dz\/2\pi iz}\left\{
{ 1\/1-{rw/z}}f^{ij}\({w\/z}\)W^i(z)W^j(w)
+{{z/rw}\/1-{z/rw}} W^j(w)W^i(z)f^{ji}\({z\/w}\)
\right\}\cr
&\!\!=\!\!&
\sum_{n\in\bZ}\sum_{m\geq 0}\sum_{\ell=0}^m f^{ij}_\ell \left\{
r^{ m-\ell}\cdot W^i_{-m} W^j_{n+m}
+r^{\ell-m-1}\cdot W^j_{n-m-1}W^i_{m+1}
\right\}w^{-n},
\end{eqnarray}
with
\begin{eqnarray}
f^{ij}(x) &\!\!\equiv\!\!& \Exp{ \sum_{n>0}{1\/n}(1-q^n)(1-t^{-n})
{1-p^{in}\/1-p^n}{1-p^{(N-j)n}\/1-p^{Nn}} p^{{j-i\/2}n} x^n },\cr
f^{ji}(x) &\!\!\equiv\!\!& f^{ij}(x),\qquad (i\leq j),
\end{eqnarray}
and
$f^{ij}(x)\equiv\sum_{\ell\geq 0}f^{ij}_\ell x^\ell$.
Here $(1-x)^{-1}$ stands for $\sum_{n\geq 0}x^n$.
Remark that this normal ordering $\:*\:$ is a generalization
of the following usual one $(*)$ used in the conformal field theory:
\begin{eqnarray}
\(AB\)(w) &\!\!\equiv\!\!& \oint_w {dz\/2\pi i} {1\/z-w} A(z) B(w)\cr
&\!\!\equiv\!\!& \oint_0 {dz\/2\pi iz}
\left\{{1\/1-{w/z}}A(z)B(w) + {{z/w}\/1-{z/w}}B(w)A(z)\right\}.
\end{eqnarray}
The relation of the {$q$-${\cal W}_N$} generators should be
written in this normal ordering.
Here we present some examples of them.
The relation of $W^1(z)$ and $W^j(z)$ for $j\geq 1$ is
\begin{eqnarray}
&\!\!\!\!\!\!&\!\!\!\!\!\!
f^{1j}\({w\/z}\)W^1(z)W^j(w) - W^j(w)W^1(z)f^{j1}\({z\/w}\) \\
&\!\!=\!\!&
-{(1-q)(1-t^{-1})\/1-p}\left\{
\,\delta\(p^{j+1\/2}{w\/z}\)W^{j+1}\(p^{1\/2}w\)
-\,\delta\(p^{-{j+1\/2}}{w\/z}\)W^{j+1}\(p^{-{1\/2}}w\)
\right\},\nonumber
\end{eqnarray}
with $\,\delta(x)\equiv \sum_{n\in\bZ} x^n$;
and that of $W^2(z)$ and $W^j(z)$ for $j\geq 2$ is
\begin{eqnarray}
&\!\!\!\!\!\!&\!\!\!\!\!\!
f^{2j}\({w\/z}\)W^2(z)W^j(w) - W^j(w)W^2(z)f^{j2}\({z\/w}\) \\
&\!\!=\!\!&
-{(1-q)(1-t^{-1})\/1-p}{(1-qp)(1-t^{-1}p)\/(1-p)(1-p^2)}
\left\{\,\delta\(p^{ {j\/2}+1}{w\/z}\)W^{j+2}(p w) \right. \cr
&\!\!\!\!\!\!&\hspace{66mm}
\left.-\,\delta\(p^{-{j\/2}-1}{w\/z}\)W^{j+2}(p^{-1}w) \right\}\cr
&\!\!\!\!\!\!&
-{(1-q)(1-t^{-1})\/1-p}
\left\{\,\delta\(p^{{j\/2}}{w\/z}\)
\: W^1(p^{-{1\/2}}z) W^{j+1}(p^{ {1\/2}} w)\: \right. \cr
&\!\!\!\!\!\!&\hspace{32mm}
\left.-\,\delta\(p^{-{j\/2}}{w\/z}\)
\: W^1(p^{ {1\/2}}z) W^{j+1}(p^{-{1\/2}} w)\: \right\}\cr
&\!\!\!\!\!\!&
+{(1-q)^2(1-t^{-1})^2\/(1-p)^2}
\left\{\,\delta\(p^{{j\/2}}{w\/z}\)
\({p^2\/1-p^2}W^{j+2}(pw)+{1\/1-p^j} W^{j+2}( w)\)\right.\cr
&\!\!\!\!\!\!&\hspace{35mm}
\left.-\,\delta\(p^{-{j\/2}}{w\/z}\)
\({p^j\/1-p^j}W^{j+2}( w)+{1\/1-p^2} W^{j+2}(p^{-1}w)\)\right\},
\nonumber
\end{eqnarray}
with $W^i(z) \equiv 0$ for $i>N$.
The main terms of
$$
f^{ij}\({w\/z}\)W^i(z)W^j(w) - W^j(w)W^i(z)f^{ji}\({z\/w}\)\qquad
(i\leq j)
$$
is
\begin{eqnarray}
&\!\!\!\!\!\!&
-{(1-q)(1-t^{-1})\/1-p}\sum_{k=1}^{{\rm min} \(i,N-j\)}
\prod_{\ell=1}^{k-1}{(1-qp^\ell)(1-t^{-1}p^\ell)\/(1-p^\ell)(1-p^{\ell+1})}\cr
&\!\!\!\!\!\!&\!\!\!\!\!\! \times
\left\{
\,\delta\(p^{{j-i\/2}+k}{w\/z}\)\:W^{i-k}(p^{-{k\/2}}z)W^{j+k}(p^{ {k\/2}}w)\:
-\,\delta\(p^{{i-j\/2}-k}{w\/z}\)\:W^{i-k}(p^{ {k\/2}}z)W^{j+k}(p^{-{k\/2}}w)\:
\right\}.
\nonumber
\end{eqnarray}
\newcommand\AC{
In these kinds of formulae we use
$ \Exp{-\sum_{n>0}x^n /n} = 1-x =
-x\Exp{-\sum_{n>0}x^{-n}/n}$.}
To obtain the above relations, the fundamental formula is
\begin{eqnarray}
f^{11}\({w\/z}\)\La_i(z)\La_i(w) &\!\!-\!\!& \La_i(w)\La_i(z)f^{11}\({z\/w}\)
= 0,\cr
f^{11}\({w\/z}\)\La_i(z)\La_j(w) &\!\!-\!\!& \La_j(w)\La_i(z)f^{11}\({z\/w}\)
\cr
&\!\!=\!\!&{(1-q)(1-t^{-1})\/1-p}\(\,\delta\({w\/z}\)-\,\delta\(p{w\/z}\)\)
:\La_i(z)\La_j(w):,
\nonumber
\end{eqnarray}
for $i<j$, here we use\footnote{\AC}
\begin{eqnarray}
&\!\!\!\!\!\!&\!\!\!\!\!\!
\Exp{ \sum_{n>0}{1\/n}(1-q^{ n})(1-t^{-n})x^{ n} }-
\Exp{ \sum_{n>0}{1\/n}(1-q^{-n})(1-t^{ n})x^{-n} }\cr
&\!\!\!\!\!\!&\!\!\!\!\!\! \hspace{55mm}
= {(1-q)(1-t^{-1})\/1-p}\(\,\delta(x)-\,\delta(px)\).
\end{eqnarray}
To calculate the general relations, the following formulae are useful:
\begin{eqnarray}
&\!\!\!\!\!\!&\!\!\!\!\!\!
\Exp{ \sum_{n>0}{1\/n}(1-q^{ n})(1-t^{-n})(1+r^{ n})x^{ n} }-
\Exp{ \sum_{n>0}{1\/n}(1-q^{-n})(1-t^{ n})(1+r^{-n})x^{-n} }\cr
&\!\!\!\!\!\!&\!\!\!\!\!\! \hspace{30mm}
={(1-q)(1-t^{-1})\/(1-p)(1-r)}
\left\{
(1-qr)(1-t^{-1}r) {\,\delta( x)-\,\delta(prx)\/1-pr} \right.\cr
&\!\!\!\!\!\!&\!\!\!\!\!\! \hspace{70mm}\left.
-(r-q )(r-t^{-1} ) {\,\delta(rx)-\,\delta(p x)\/r-p }
\right\},
\label{e:formula.3}
\end{eqnarray}
with $r\neq 0$;
For $r=1$ or $p^{\pm1}$,
the right hand side of \eq{e:formula.3} should be understood as the limit
$r\rightarrow 1$ or $p^{\pm1}$, respectively;
And
$f^{ij}(x) = \prod_{k=1}^i f^{1j}(p^{{i+1\/2}-k}x)$
for $i\leq j$.
\subsection{Example of $q$-$\cW_3$}
$N=2$ case is {${\cal V}ir\hspace{-.03in}_{q,t}\,$} studied in Ref. \cite{rSKAO} (see appendix A).\\
Here we give an example when $N=3$.
The generators are
\begin{eqnarray}
W^1(z) &\!\!=\!\!& \La_1(z) + \La_2(z) + \La_3(z),\cr
W^2(z) &\!\!=\!\!&
\La_1(zp^{1\/2})\La_2(zp^{-{1\/2}}) +
\La_1(zp^{1\/2})\La_3(zp^{-{1\/2}}) +
\La_2(zp^{1\/2})\La_3(zp^{-{1\/2}}).
\end{eqnarray}
The relation of these generators is
\begin{eqnarray}
\&\hspace{-5mm}
f^{11}\({w\/z}\) W^1(z) W^1(w) - W^1(w) W^1(z) f^{11}\({z\/w}\) \cr
\&\hspace{10mm}
= -{(1-q)(1-t^{-1})\/1-p}\left\{
\,\delta\({w\/z}p \) W^2\(wp^{ 1\/2 }\)-
\,\delta\({w\/z}p^{-1}\) W^2\(wp^{-{1\/2}}\)\right\},\cr
\&\hspace{-5mm}
f^{12}\({w\/z}\) W^1(z) W^2(w) - W^2(w) W^1(z) f^{21}\({z\/w}\) \cr
\&\hspace{10mm}
= -{(1-q)(1-t^{-1})\/1-p}\left\{
\,\delta\({w\/z}p^{ {3\/2}}\)-
\,\delta\({w\/z}p^{-{3\/2}}\)\right\},\cr
\&\hspace{-5mm}
f^{22}\({w\/z}\) W^2(z) W^2(w) - W^2(w) W^2(z) f^{22}\({z\/w}\) \cr
\&\hspace{10mm}
= -{(1-q)(1-t^{-1})\/1-p}\left\{
\,\delta\({w\/z}p \) W^1\(zp^{-{1\/2}}\)-
\,\delta\({w\/z}p^{-1}\) W^1\(zp^{ 1\/2 }\)\right\},\nonumber
\end{eqnarray}
with
\begin{eqnarray}
f^{11}(x) &\!\!=\!\!&
\Exp{ \sum{1\/n}(1-q^n)(1-t^{-n}){1-p^{2n}\/1-p^{3n}}x^n }
=f^{22}(x),\cr
f^{12}(x) &\!\!=\!\!&
\Exp{ \sum{1\/n}(1-q^n)(1-t^{-n}){1-p^n\/1-p^{3n}}p^{n\/2}x^n }
=f^{21}(x).\nonumber
\end{eqnarray}
Note that there is no difference between $W^1$ and $W^2$
in algebraically.
\subsection{Highest weight module of {$q$-${\cal W}_N$} algebra}
Here we refer to the representation of the {$q$-${\cal W}_N$} algebra.
Let $|\la\>$ be the highest weight vector of the {$q$-${\cal W}_N$} algebra
which satisfies
$W^i_n|\la\>=0$ for $n>0$ and $i=1,2,\cdots,N-1$ and
$W^i_0|\la\>= \la^i |\la\>$ with $\la^i\in\bC$.
Let $M_\la$ be the Verma module over the {$q$-${\cal W}_N$} algebra
generated by $|\la\>$.
The dual module $M_\la^*$ is generated by $\<\la|$ such that
$\<\la|W^i_n=0$ for $n<0$ and $\<\la|W^i_0= \la^i\<\la|$.
The bilinear form $M_\la^*\otimes M_\la\rightarrow\bC$
is uniquely defined by $\<\la|\la\>=1$.
A singular vector $|\chi\>\in M_\la$ is defined by
$W^i_n|\chi\>=0$ for $n>0$ and
$W^i_0|\chi\>= (\la^i+N^i) |\chi\>$
with $N^i\in\bC$.
\section{Screening currents and singular vectors}
Next we turn to the screening currents, a commutant of the {$q$-${\cal W}_N$} algebra,
which construct the singular vectors.
\subsection{Screening currents}
Let us introduce root bosons
${\al in} \equiv {\h in}-{\h {i+1}n}$ and
${\Qal i} \equiv {\Qh i}-{\Qh {i+1}}$ for $i=1,2,\cdots,N-1$.
Then they satisfies
\begin{eqnarray}
[{\al in},{\al jm}] &\!\!=\!\!& -{1\/n}(1-q^n)(1-t^{-n})
\left\{(1+p^{-n})\,\delta_{i,j} - \,\delta_{i+1,j} -
p^{-n}\,\delta_{i-1,j}\right\}
\,\delta_{n+m,0},\cr
[{\al i0},{\Qal j}] &\!\!=\!\!& 2\,\delta_{i,j}-\,\delta_{i+1,j}-\,\delta_{i-1,j},
\end{eqnarray}
and
\begin{eqnarray}
[{\h in},{\al jm}] &\!\!=\!\!& {1\/n}(1-q^{-n})(1-t^{-n})
\left\{ q^n\,\delta_{i,j} - t^n\,\delta_{i,j+1}\right\}
\,\delta_{n+m,0},\cr
[{\h i0},{\Qal j}] &\!\!=\!\!& \,\delta_{i,j}-\,\delta_{i,j+1},\qquad
[{\al i0},{\Qh j}] = \,\delta_{i,j}-\,\delta_{i+1,j}.
\end{eqnarray}
Note that
$[{\h in} + p^n {\h {i+1}n}, {\al im}] = 0$.
By using these root bosons, we define screening currents as follows:
\begin{eqnarray}
S^i_+(z) &\!\!\equiv\!\!&
:\Exp{ \sum_{n\neq0}{{\al in}\/1-q^n} z^{-n} }:
e^{\rb{\Qal i}} z^{\rb{\al i0}},\cr
S^i_-(z) &\!\!\equiv\!\!&
:\Exp{ -\sum_{n\neq0}{{\al in}\/1-t^n} z^{-n} }:
e^{\rbi{\Qal i}} z^{\rbi{\al i0}}.
\end{eqnarray}
Then we have
\proclaim Proposition.
The screening currents satisfy
\begin{eqnarray}
\&\hspace{-7mm}
\[\,:\!\(p^{D_z} - \La_1(z)\)\(p^{D_z} - \La_2(zp^{-1})\)\cdots
\(p^{D_z} - \La_N(zp^{1-N})\)\!:\,,
S^i_\pm(w)\]\cr
\&
= (1-q^{\pm1})(1-t^{\mp1}){d\/d_{q\atop t}w}
\,:\!\(p^{D_z} - \La_1(z)\)\cdots\(p^{D_z} - \La_{i-1}(zp^{2-i})\) \cr
\&\hspace{5mm}\times
w \,\delta\({w\/z}p^{i-1}\) A^i_\pm(w) p^{D_z}
\(p^{D_z} - \La_{i+2}(zp^{-1-i})\)\cdots\(p^{D_z} - \La_N(zp^{1-N})\) \!:\,,
\nonumber
\end{eqnarray}
with
\begin{eqnarray}
A^i_+(w) &\!\!=\!\!&
:\Exp{ \sum_{n\neq0}{{\h in}-q^n{\h {i+1}n}\/1-q^n}w^{-n} }:\,
e^{\rb{\Qal i}} w^{\rb{\al i0}} q^{\rb{\h{i+1}0}} p^{{N+1\/2}-i-1},\cr
A^i_-(w) &\!\!=\!\!&
:\Exp{ -\sum_{n\neq0}{t^n{\h in}-{\h {i+1}n}\/1-t^n}w^{-n} }:\,
e^{\rbi{\Qal i}} w^{\rbi{\al i0}} q^{\rb{\h i0}} p^{{N+1\/2}-i}.\nonumber
\end{eqnarray}
\noindent
Here ${d\/d_\xi w} f(w) \equiv (f(w)-f(\xi w))/((1-\xi)w)$.
\noindent{\it Proof.\quad}
First, we have
\begin{eqnarray}
[\La_i(z),S^j_+(w)] &\!\!=\!\!&
(t-1)\,\delta_{i,j}\,\delta\({w\/z}q\):\La_j(z) S^j_+(w):\cr
&&+
(t^{-1}-1)\,\delta_{i,j+1}\,\delta\({w\/z}\):\La_{j+1}(z) S^j_+(w):,\cr
[\La_i(z),S^j_-(w)] &\!\!=\!\!&
(q^{-1}-1)\,\delta_{i,j}\,\delta\({w\/z}\):\La_j(z) S^j_-(w):\cr
&&+
(q-1)\,\delta_{i,j+1}\,\delta\({w\/z}t\):\La_{j+1}(z) S^j_-(w):.
\end{eqnarray}
Here we use the following formula:
\begin{equation}
q^{\mp1}
\Exp{ \pm\sum_{n>0}{1\/n}(1-q^n )x^n } -
\Exp{ \pm\sum_{n>0}{1\/n}(1-q^{-n})x^{-n} } =
(q^{\mp1}-1)\,\delta\(xq^{1\mp1\/2}\).
\end{equation}
The operator parts are
\begin{eqnarray}
:\La_j(wq) S^j_+(w): &\!\!=\!\!& A^j_+(wq) p,\qquad
:\La_{j+1}(w) S^j_+(w):\,= A^j_+(w),\cr
:\La_j(w) S^j_-(w): &\!\!=\!\!& A^j_-(w),\qquad
:\La_{j+1}(wt) S^j_-(w):\,= A^j_-(wt) p^{-1}.
\end{eqnarray}
Next,
\begin{eqnarray}
[\La_i(z)+\La_{i+1}(z), S^i_\pm(w)] &\!\!=\!\!& -(1-q^{\pm1})(1-t^{\mp1})
{d\/d_{q\atop t}w}\left\{w\,\delta\({w\/z}\)A^i_\pm(w)\right\},\cr
[:\La_i(z)\La_{i+1}(zp^{-1}):, S^i_\pm(w)] &\!\!=\!\!& 0.
\end{eqnarray}
Hence,
\begin{eqnarray}
\&
\[\,:\!\(p^{D_z} - \La_i(z)\)\(p^{D_z} - \La_{i+1}(zp^{-1})\)\!:\,,
S^i_\pm(w)\]\cr
\&\hspace{35mm}
= (1-q^{\pm1})(1-t^{\mp1})
{d\/d_{q\atop t}w}\left\{w\,\delta\({w\/z}\)A^i_\pm(w)\right\} p^{D_z}.
\end{eqnarray}
This gives us the proposition. \hfill\fbox{}
Therefore,
the screening currents $S^i_\pm(z)$
commute with any {$q$-${\cal W}_N$} generators up to total difference.
Thus we obtain
\proclaim Theorem.
Screening charges $\oint dz S^i_\pm(z)$
commute with any {$q$-${\cal W}_N$} generators.
\subsection{Singular vectors}
Let $\cF_\a$ be the boson Fock space
generated by the highest weight state $|\a\>$ such that
${\al in} |0\> = 0$ for $n\geq0$ and
$|\a\> \equiv \exp\{\sum_{i=1}^{N-1}\a^i{\QLa i}\} |0\>$
with ${\QLa i}\equiv\sum_{j=1}^i {\Qh j}$.
Note that ${\al i0}|\a\> = \a^i|\a\>$.
And this state $|\a\>$ is also the highest weight state of the {$q$-${\cal W}_N$} algebra.
We denote the negative mode part of $S^i_+(z)$ as
$(S^i_+(z))_- \equiv \Exp{ \sum_{n<0}{{\al in}\/1-q^n} z^{-n} }$.
Then we have
\proclaim Proposition.
For a set of non-negative integers $s_a$ and $r_a\geq r_{a+1}\geq0$,
($a=1,\cdots,N-1$),
let
\begin{eqnarray}
\a_{r,s}^{a} &\!\!=\!\!&\rb(1+r_a-r_{a-1})\rbi(1+s_a),\qquad
r_0 = 0,\cr
\widetilde\a_{r,s}^{a} &\!\!=\!\!&\rb(1-r_a+r_{a+1})\rbi(1+s_a),\qquad
r_N = 0.
\label{e:weightalpha}
\end{eqnarray}
Then the singular vectors $|\chi_{rs}^+\>\in\cF_{\a_{rs}^+}$
are realized by the screening currents as follows:
\begin{eqnarray}
\&\hspace{-5mm}
|\chi_{r,s}\> =
\oint\prod_{a=1}^{N-1}\prod_{j=1}^{r_a} {dx^a_j}\cdot
S^1_+(x^1_1)\cdots S^1_+(x^1_{r_1}) \cdots
S^{N-1}_+(x^{N-1}_1) \cdots S^{N-1}_+(x^{N-1}_{r_{N-1}})
|\widetilde\a_{r,s}\>\cr
&\!\!=\!\!&
\oint\prod_{a=1}^{N-1} \prod_{j=1}^{r_a} {dx^a_j\/x^a_j}\cdot
\prod_{a=1}^{N-1} \Pi\(\overline{x^a},px^{a+1}\) \Delta(x^a) C(x^a)
\prod_{j=1}^{r_a} (x^a_j)^{-s_a} (S^a_+(x^a_j))_-\cdot|\a_{r,s}\>\cr
&&\label{e:singular}
\end{eqnarray}
with $x^N=0$, $\overline x = 1/x$ and
\begin{eqnarray}
\Pi(x,y) &\!\!=\!\!& \prod_{ij}
\Exp{ \sum_{n>0}{1\/n}{1-t^n\/1-q^n} x_i^n y_j^n },\qquad
\Delta(x) =
\prod_{i\neq j}^r\Exp{ -\sum_{n>0}{1\/n}{1-t^n\/1-q^n}{x_j^n\/x_i^n}},\cr
C(x) &\!\!=\!\!& \prod_{i<j}^r
\Exp{\sum_{n>0}{1\/n}{1-t^n\/1-q^n}\({x_i^n\/x_j^n}-p^n{x_j^n\/x_i^n}\)}
\prod_{i=1}^r x_i^{(r+1-2i)\beta}
\label{e:Delta}
\end{eqnarray}
\noindent{\it Proof.\quad}
The operator product expansion of the screening currents is
\begin{eqnarray}
S^a_+(x) S^a_+(y) &\!\!=\!\!&
\Exp{-\sum_{n>0}{1\/n}{1-t^n\/1-q^n}(1+p^n){y^n\/x^n}}x^{2\beta}
:S^a_+(x) S^a_+(y):,\cr
S^a_+(x) S^{a\pm1}_+(y) &\!\!=\!\!&
\Exp{\sum_{n>0}{1\/n}{1-t^n\/1-q^n}p^{{1\pm1\/2}n}{y^n\/x^n}}x^{-\beta}
:S^a_+(x) S^{a\pm1}_+(y):.
\end{eqnarray}
Since
\begin{eqnarray}
S^a_+(x_1)\cdots S^a_+(x_r)
&\!\!=\!\!&
\prod_{i<j}
\Exp{-\sum_{n>0}{1\/n}{1-t^n\/1-q^n}(1+p^n){x_j^n\/x_i^n}}
\prod_{i=1}^r x_a^{2\beta(r-i)}
:\prod_{i=1}^r S^a_+(x_i):\cr
&\!\!=\!\!&
\Delta(x) C(x)\prod_{i=1}^r x_i^{(r-1)\beta}:\prod_{i=1}^r S^a_+(x_i):,
\end{eqnarray}
and
\begin{equation}
:\prod_{a=1}^{N-1}\prod_{i=1}^{r_a} S^a_+(x_i):|\widetilde\a_{r,s}\>
=\prod_{a=1}^{N-1}\prod_{i=1}^{r_a}
(x^a_i)^{(1-r_a+r_{a+1})\beta-(1+s_a)}(S^a_+(x_i))_- \cdot|\a_{r,s}\>,
\end{equation}
we obtain the proposition. \hfill\fbox{}
Note that $C(x)$ is a pseudo-constant under the $q$-shift, {\it i.e.,}
$q^{D_{x_i}}C(x)=C(x)$.
The expression in \eq{e:weightalpha}
is the same as that of $q=1$ case \cite{rAMOS}.
Remark that the singular vectors are also realized
by using the other screening currents $S_-^i(x)$
by the replacing $t$ with $q^{-1}$ and $\rb$ with $-1/\rb$ in
\eq{e:singular}, that is to say:
\begin{eqnarray}
\&\hspace{-5mm}
|\chi_{r,s}^-\> =
\oint\prod_{a=1}^{N-1}\prod_{j=1}^{r_a} {dx^a_j}\cdot
S^1_-(x^1_1)\cdots S^1_-(x^1_{r_1}) \cdots
S^{N-1}_-(x^{N-1}_1) \cdots S^{N-1}_-(x^{N-1}_{r_{N-1}})
|\widetilde\a_{r,s}^-\>\cr
&\!\!=\!\!&
\oint\prod_{a=1}^{N-1} \prod_{j=1}^{r_a} {dx^a_j\/x^a_j}\cdot
\prod_{a=1}^{N-1} \Pi_-\(\overline{x^a},x^{a+1}\) \Delta_-(x^a) C_-(x^a)
\prod_{j=1}^{r_a} (x^a_j)^{-s_a} (S^a_-(x^a_j))_-\cdot|\a_{r,s}^-\>,\cr
&&\label{e:singularMinus}
\end{eqnarray}
where $\widetilde\a_{r,s}^-$, $\a_{r,s}^-$, $\Pi_-$, $\Delta_-$ and $C_-$
are obtained from those without $-$ suffix
by the replacing $t$ with $q$ and $\rb$ with $-1/\rb$.
And $(S^a_-(z))_-$ is the negative mode part of $S^a_-(z)$.
\section{Vertex operator of fundamental representation}
Now we introduce a vertex operator.
Let $V(z)$ be the vertex operator defined as
\begin{equation}
V(z) \equiv
\,:\!\Exp{ -\sum_{n\neq0}{{\h 1n}\/1-q^n} p^{-{n\/2}}z^{-n} }\!:\,
e^{-\rb{\Qh 1}} z^{-\rb{\h 10}}.
\label{e:vertex}
\end{equation}
When $q=1$,
this $V(z)$ coincides with the vertex operator of fundamental representation.
Note that the fundamental vertex $\La_1(z)$ can be realized
by $V(z)$ as
\begin{equation}
\La_1(zp^{1\/2}) = \,:\!V(zq^{-1}) V^{-1}(z)\!:\, p^{N-1\/2}.
\end{equation}
Hence, this vertex operator $V(z)$ can be considered as
one of a building block of the {$q$-${\cal W}_N$} generators.
We have
\proclaim Proposition.
The vertex operator $V(w)$ enjoys the following Miura-like relation:
\begin{eqnarray}
\&\hspace{-5mm}
:\!\(p^{D_z} - g^L\({w\/z }\) \La_1(z )\)\cdots
\(p^{D_z} - g^L\({w\/zp^{1-N}}\) \La_N(zp^{1-N})\)\!:
V(w)\cr
\&
- V(w)
:\!\(p^{D_z} - \La_1(z ) g^R\({z \/w}\)\)\cdots
\(p^{D_z} - \La_N(zp^{1-N}) g^R\({zp^{1-N}\/w}\)\)\!:
\cr
\& \hspace{5mm}
= p^{N-1\/2}(1-t^{-1})\,\delta\({w\/z}p^{1\/2}\)
:V(wq^{-1}) \(p^{D_z} - \La_2(zp^{-1})\)\cdots\(p^{D_z} - \La_N(zp^{1-N})\):,
\nonumber
\end{eqnarray}
and
\begin{eqnarray}
g^L(x) &\!\!=\!\!&
\Exp{\sum_{n>0} {1\/n}(1-t^{ n}){1-p^n \/1-p^{ Nn}}p^{ {n\/2}}x^n }
t^{-{1\/N}},\cr
g^R(x) &\!\!=\!\!&
\Exp{\sum_{n>0} {1\/n}(1-t^{-n}){1-p^{-n}\/1-p^{-Nn}}p^{-{n\/2}}x^n }.
\end{eqnarray}
\noindent{\it Proof.\quad}
The fundamental relation is
\begin{equation}
g^L\({w\/z}\) \La_i(z) V(w) - V(w) \La_i(z) g^R\({z\/w}\) =
p^{{N-1\/2}} (t^{-1}-1) \,\delta_{i,1} \,\delta\({w\/z}p^{1\/2}\) V(wq^{-1}),
\end{equation}
{\it i.e.,}
\begin{eqnarray}
\(p^{D_z} - g^L\({w\/z}\) \La_i(z)\)V(w)
&\!\!=\!\!&
V(w)\(p^{D_z} - \La_i(z) g^R\({z\/w}\)\)\cr
&\!\!+\!\!&
p^{N-1\/2}(1-t^{-1})\,\delta_{i,1}\,\delta\({w\/z}p^{1\/2}\) V(wq^{-1}),
\label{e:LaV}
\end{eqnarray}
here we use
$:\!\La_1(wp^{1\/2}) V(w)\!:\, = V(wq^{-1}) p^{{N-1\/2}}$.
By using this relation \eq{e:LaV} and
$V(w) \La_i(z) g^R\({z/w}\) = :\!V(w)\La_i(z)\!:$,
we obtain the proposition.
\hfill\fbox{}
For example, when $N=3$,
the relation between the vertex operator $V(w)$ and the {$q$-${\cal W}_N$} generators is
\begin{eqnarray}
\&
g^L\({w\/z}\) W^1(z) V(w) - V(w) W^1(z) g^R\({z\/w}\)
=p (t^{-1}-1)\,\delta\({w\/z}p^{1\/2}\) V(wq^{-1}),\cr
\&
g^L\({w\/z}\) g^L\({w\/z}p\) W^2(zp^{-{1\/2}}) V(w) -
V(w) W^2(zp^{-{1\/2}}) g^R\({z\/w}\) g^R\({z\/w}p^{-1}\) \cr
\&\hspace{5mm}
=p (t^{-1}-1)\,\delta\({w\/z}p^{1\/2}\)
\(\,:\!V(wq^{-1})\La_2(wp^{-{1\/2}})\!:\,+
\,:\!V(wq^{-1})\La_3(wp^{-{1\/2}})\!:\,\).
\end{eqnarray}
\section{Macdonald polynomials}
Finally we present a relation with the Macdonald polynomials.
The excited states of trigonometric Ruijsenaars model are
called Macdonald symmetric functions $P_\la(z)$
and they are defined as follows:
\begin{eqnarray}
&&\qquad
H P_\la(z_1,\cdots,z_M) =\varepsilon_\la P_\la(z_1,\cdots,z_M),\cr
&&
H = \sum_{i=1}^M \prod_{j\neq i}
{t z_i - z_j \/ z_i - z_j
\cdot q^{D_{z_i}},\qquad
\varepsilon_\la = \sum_{i=1}^M t^{M-i} q^{\la_i},
\label{e:macDef}
\end{eqnarray}
where
the $\la = (\la_1\geq\la_2\geq\cdots\la_M\geq0)$ is a partition.
The Macdonald polynomials with general Young diagram $\la$
are realized as some kinds of correlation functions of
the screening currents and vertex operators of the {$q$-${\cal W}_N$} algebra
as follows:
\proclaim Theorem.
Macdonald polynomial $P_\la(z)$ with the Young diagram
$\la = \sum_{i=1}^{N-1} (s_i^{r_i})$, $r_i\geq r_{i+1}$
is written as
\begin{equation}
P_\la\(z_1,\cdots,z_M\)\propto
\<\a_{r,s}|\Exp{-\sum_{n>0}{{\h 1n}\/1-q^n}\sum_{i=1}^Mz_i^n}|\chi_{r,s}\>.
\end{equation}
Here $|\chi_{r,s}\>$ is a singular vector in \eq{e:singular}.
Note that the operator part of the above equation
is the positive mode part of the product of the vertex operators \eq{e:vertex}.
The Young diagram is as follows:
\generalYoung
\noindent{\it Proof.\quad}
First we have
\begin{equation}
\Exp{ -\sum_{n>0}{{\h 1n}\/1-q^n}\sum_{i=1}^M z_i^n} S^a_+(w) =
\Pi\(z,px^1\)^{\,\delta_{a,1}}
S^a_+(w) \Exp{ -\sum_{n>0}{{\h 1n}\/1-q^n}\sum_{i=1}^M z_i^n}.
\end{equation}
By \eq{e:singular},
the right hand side of the equation of this theorem is
\begin{equation}
\oint\prod_{a=1}^{N-1} \prod_{j=1}^{r_a} {dx^a_j\/x^a_j}\cdot
\Pi\(z,px^1\)
\prod_{a=1}^{N-1} \Pi\(\overline{x^a},px^{a+1}\) \Delta(x^a) C(x^a)
\prod_{j=1}^{r_a} (x^a_j)^{-s_a},
\label{e:macOPE}
\end{equation}
If we replace $x^a$ with $(p^a x^a)^{-1}$ in \eq{e:macOPE},
then the integrand coincides with that of the integral formula for
Macdonald polynomials in Ref. \cite{rAOS}
except for the $C(x)$ parts.
For the integral representation of the Macdonald polynomial,
we need only the property with respect to a $q$-shift.
Since this $C(x)$ is a pseudo-constant under it, {\it i.e.,}
$q^{D_{x_i}}C(x)=C(x)$,
they are integral representations of the Macdonald polynomial
(see appendix B).
\hfill\fbox{}
Remark that the Macdonald polynomials with the dual Young diagram
$\la'= \(r_1^{s_1},r_2^{s_2},\cdots,r_{N-1}^{s_{N-1}}\)$
are realized by using the other screening currents $S_-^i(x)$
with $|\chi_{r,s}^-\>$ in \eq{e:singularMinus} as
\begin{equation}
P_{\la'}\(-z\)\propto
\<\a_{r,s}^-|\Exp{-\sum_{n>0}{{\h 1n}\/1-q^n}\sum_{i=1}^Mz_i^n}|\chi_{r,s}^-\>.
\end{equation}
\section{Conclusion and discussion}
\def\FeiginFrenkel{
After finishing of this work,
we received the preprint
{\it ``Quantum $\cW$-algebras and elliptic algebras''}
by B.~Feigin and E.~Frenkel (q-alg/9508009).
They discuss similar things with ours.
Although the algebra of screening currents is considered there,
the normal ordering of $q$-$\cW$ generators and
the relation with the Macdonald polynomial are not given.
}
We have derived a quantum $\cW_N$ algebra
whose some kinds of correlation functions are the Macdonald polynomials.
\footnote\FeiginFrenkel
Jack polynomials are realized in the following two ways
(see also \cite{rLV}):
one is some kinds of correlation function of $\cW_N$ algebra
\cite{rMY,rAMOS},
the other is suitable combinations of
correlation functions of $\widehat{sl_N}$ algebra \cite{rMC}.
The relations between Macdonald polynomials,
the {$q$-${\cal W}_N$} algebra and the $U_q(\widehat{sl_N})$ algebra
are interesting.
In the classical limit $\hbar\rightarrow 0$ with $q\equiv e^\hbar$,
$q$-Miura transformation \eq{e:qMiura} reduces to
the classical one.
Since the right hand side of it is order $\hbar^N$,
the left hand side must be the same order.
To do so, $\hbar$ expansion of the {$q$-${\cal W}_N$} generators must be nontrivial.
Moreover, the classical generators are obtained as a linear
combination of the {$q$-${\cal W}_N$} generators.
\vskip5mm
\noindent{\bf Acknowledgments:}
\noindent
We would like to thank
B.~Feigin, E.~Frenkel and Y.~Matsuo
for valuable discussions.
S.O. would like to thank members of YITP for their hospitality.
This work is supported in part by Grant-in-Aid for Scientific
Research from Ministry of Science and Culture.
\section*{Appendix A: Quantum Virasoro algebra}
\def\Lukyanov{
The same operator with $S^1_+(z)$ was considered in \cite{rPL}.
}
In this appendix, we give an example when $N=2$,
{\it i.e.,} {${\cal V}ir\hspace{-.03in}_{q,t}\,$} in \cite{rSKAO}.
The fundamental bosons ${\h 1n}$ and ${\Qh 1}$ satisfy
\begin{equation}
[{\h 1n},{\h 1m}] = -{1\/n}{(1-q^n)(1-t^{-n})\/1+p^n}\,\delta_{n+m,0},\qquad
[{\h 10},{\Qh 1}] = {1\/2}.
\end{equation}
The root bosons are
${\al 1n} = (1+p^{-n}) {\h 1n}$ and ${\Qal 1} = 2 {\Qh 1}$.
The $q$-Virasoro generator $W^1(z)$,
the screening currents $S^1_\pm(z)$ and
the vertex operator $V(z)$ are now\footnote\Lukyanov
\begin{eqnarray}
W^1(z) &\!\!=\!\!&
:\Exp{ \sum_{n\neq 0}{\h 1n} z^{-n}}:q^{ \rb{\h 10}} p^{ 1\/2 }+
:\Exp{-\sum_{n\neq 0}{\h 1n}p^{-n}z^{-n}}:q^{-\rb{\h 10}} p^{-{1\/2}},\cr
S^1_\pm(z) &\!\!=\!\!&
:\Exp{\pm\sum_{n\neq0}{1+p^{-n}\/1-r_\pm^n}{\h 1n} z^{-n}}:
e^{\pm2\rb^{\pm1}{\Qh 1}} z^{\pm2\rb^{\pm1}{\h 10}},\quad
r_+ = q,\quad r_- = t,\cr
V(z) &\!\!=\!\!&
\,:\Exp{ -\sum_{n\neq0}{{\h 1n}\/1-q^n} p^{-{n\/2}}z^{-n} }:\,
e^{-\rb{\Qh 1}} z^{-\rb{\h 10}}.
\end{eqnarray}
The relations of them are
\begin{eqnarray}
f^{11}\({w\/z}\) W^1(z) W^1(w) &\!\!-\!\!& W^1(w) W^1(z) f^{11}\({z\/w}\) \cr
&\!\!=\!\!& -{(1-q)(1-t^{-1})\/1-p}
\left\{\,\delta\({w\/z}p\)-\,\delta\({w\/z}p^{-1}\)\right\},\cr
f^{11}(x) &\!\!=\!\!& \Exp{\sum_{n>0}{1\/n}{(1-q^n)(1-t^{-n})\/1+p^n}x^n}
\end{eqnarray}
\begin{eqnarray}
\[\,W^1(z), S^1_\pm(w)\]
&\!\!=\!\!& -(1-q^{\pm1})(1-t^{\mp1})
{d\/d_{r_\pm}w}
\left\{w\,\delta\({w\/z}\)A^1_\pm(w)\right\},\cr
A^1_\pm(w) &\!\!=\!\!&
:\Exp{\sum_{n\neq0}{1+r_\mp^{\pm n}\/1-r_\pm^{\pm n}}{\h 1n}w^{-n} }:\,
e^{\pm2\rb^{\pm1}{\Qh 1}}w^{\pm2\rb^{\pm1}{\h 10}}
q^{\mp\rb{\h 10}}p^{\mp{1\/2}},\nonumbe
\end{eqnarray}
\begin{eqnarray}
g^L\({w\/z}\) W^1(z) V(w) &\!\!-\!\!& V(w) W^1(z) g^R\({z\/w}\)
= p^{1\/2}(t^{-1}-1) \,\delta\({w\/z}p^{1\/2}\) V(wq^{-1}),\cr
g^{L\atop R}(x) &\!\!=\!\!&
\Exp{\sum_{n>0} {1\/n}{1-t^{\pm n}\/1+p^{\pm n}}p^{\pm{n\/2}}x^n }
t^{-{1\pm1\/4}}.
\end{eqnarray}
For non-negative integers $s$ and $r\geq0$,
the singular vectors $|\chi_{rs}\>\in\cF_{\a_{rs}}$ are
\begin{eqnarray}
|\chi_{r,s}\> &\!\!=\!\!&
\oint\prod_{j=1}^{r} {dx_j}\cdot
S^1_+(x_1)\cdots S^1_+(x_{r}) |\a_{-r,s}\>\cr
&\!\!=\!\!&
\oint\prod_{j=1}^{r} {dx_j\/x_j}\cdot
\Delta(x) C(x)\prod_{j=1}^{r} (x_j)^{-s} (S_+(x_j))_-\cdot|\a_{r,s}\>
\end{eqnarray}
with
$\a_{r,s}^1 =\rb(1+r)\rbi(1+s)$.
$\Delta(x)$ and $C(x)$ are the same as \eq{e:Delta}.
\section*{Appendix B: Integral formula for the Macdonald polynomials}
Finally, we recapitulate the integral representation
of the Macdonald polynomials \cite{rAOS}
(\cite{rMY2,rAMOS} in the $q=1$ case).
Let us denote the Macdonald polynomial defined by \eq{e:macDef} as
$P_\la(z;q,t)$ or $P_\la(z_1,\cdots,z_M;q,t)$.
\proclaim Proposition.
The Macdonald polynomials with the Young diagram
$\la = \sum_{i=1}^{N-1} \(s_i^{r_i}\)$ or with its dual
$\la'= \(r_1^{s_1},r_2^{s_2},\cdots,r_{N-1}^{s_{N-1}}\)$ are
realized as follows:
\begin{eqnarray}
P_\la(z;q,t)&\!\!\propto\!\!&
\oint\prod_{a=1}^{N-1} \prod_{j=1}^{r_a} {dx^a_j\/x^a_j}\cdot
\Pi\(z,\overline{x^1}\)
\prod_{a=1}^{N-1} \Pi\(x^a,\overline{x^{a+1}}\) \Delta(x^a) C(x^a)
\prod_{j=1}^{r_a} (x^a_j)^{s_a},\cr
P_{\la'}(z;t,q)&\!\!\propto\!\!&
\oint\prod_{a=1}^{N-1} \prod_{j=1}^{r_a} {dx^a_j\/x^a_j}\cdot
\widetilde\Pi\(z,\overline{x^1}\)
\prod_{a=1}^{N-1} \Pi\(x^a,\overline{x^{a+1}}\) \Delta(x^a) C(x^a)
\prod_{j=1}^{r_a} (x^a_j)^{s_a},\nonumber
\end{eqnarray}
with an arbitrary pseudo-constant $C(x)$ such that $q^{D_{x_i}}C(x)=C(x)$.
Here $\widetilde\Pi(x,y)\equiv \prod_{ij}(1+x_i y_j)$.
$\Pi$ and $\Delta$ are in \eq{e:Delta}.
\noindent{\it Proof.\quad}
This proposition is proved by using
two transformations in the following lemmas iteratively.
The first transformation adds a rectangle to the Young diagram
and the second one increases the number of variables.
\hfill\fbox{}
\proclaim Lemma 1. Galilean transformation.
$($eq.\ $(VI.4.17)$ in $\cite{rM})$
\begin{equation}
P_{\la+(s^r)}(x_1,\cdots,x_r) = P_\la(x_1,\cdots,x_r)\prod_{i=1}^r x_i^s.
\end{equation}
This transformation adds a rectangle Young diagram to the original one:
$$
\Galilei
$$
\proclaim Lemma 2. Particle number changing transformation.
\begin{eqnarray}
P_\la(x_1,\cdots,x_N;q,t) &\!\!\propto\!\!&
\oint \prod_{j=1}^M {dy_j\/y_j}
\Pi(x,\overline y)\Delta(y) C(y) P_\la(y_1,\cdots,y_M;q,t),\cr
P_{\la'}(x_1,\cdots,x_N;t,q) &\!\!\propto\!\!&
\oint \prod_{j=1}^M {dy_j\/y_j}
\widetilde\Pi(x,\overline y)\Delta(y) C(y) P_\la(y_1,\cdots,y_M;q,t),\nonumber
\end{eqnarray}
here $C(y)$ is an arbitrary pseudo-constant $q^{D_{y_i}}C(y)=C(y)$
and $\la'$ is a dual Young diagram of $\la$.
\noindent{\it Proof.\quad}
Let us define scalar products $\<*,*\>$ and the another one $\<*,*\>'_N$
as follows:
\begin{eqnarray}
\<f,g\> &\!\!\equiv\!\!&
\oint\prod_{n>0} {dp_n\/2\pi i p_n}\,{f(\overline p)}\,g(p),\cr
\<f,g\>'_N &\!\!\equiv\!\!& {1\/N!}
\oint\prod_{j=1}^N {dx_j\/2\pi i x_j}\Delta(x)\,{f(\overline x)}\,g(x),
\end{eqnarray}
for the symmetric functions $f$ and $g$
with $p_n\equiv\sum_{i=1}^N x_i^N$,
$\overline{p_n}\equiv n{1-q^n\/1-t^n}{\partial\/\partial p_n}$ and
$\overline{x_j}\equiv{1/x_j}$ .
Here we must treat the power-sums $p_n$ as formally independent variables,
{\it i.e.},
${\partial\/\partial p_n}\, p_m = \delta_{n,m}$ for all $n,m>0$.
Then (eq.\ (VI.4.13) and (VI.5.4) in \cite{rM})
\begin{eqnarray}
\Pi(x,y) &\!\!=\!\!&
\sum_\la P_\la(x;q,t) P_\la(y;q,t) \<P_\la,P_\la\>^{-1},\cr
\widetilde\Pi(x,y) &\!\!=\!\!&
\sum_\la P_\la(x;q,t) P_{\la'}(y;t,q).
\label{e:completeness}\end{eqnarray}
Since the Macdonald operator is self-adjoint
for the another scalar product $\<*,*\>'_N$, that is to say
$\<H\,f,g\>'_N = \<f,H\,g\>'_N$ (eq.\ (VI.9.4) in \cite{rM}),
the Macdonald polynomials are orthogonal for this product
$\<P_\la,C\,P_\mu\>'_N \propto \delta_{\la,\mu}$
with an arbitrary pseudo-constant $C$.
The proposition follows from the completeness \eq{e:completeness} and
the orthogonality of $P_\la$'s.
\hfill\fbox{}
Remark that the above lemma 2 is also proved directly by using
the power-sum representation of the Macdonald operator \cite{rAMOS}.
Since that is also important to analyze the algebraic properties of
the Macdonald polynomials, we review it here.
\proclaim Proposition.
Macdonald operator $H(x_1,\cdots,x_N)$ are written
by the power sums $p_n \equiv \sum_{i=1}^N x_i^n$ as follows:
\begin{equation}
H = {t^N\/t-1}\oint{d\xi\/2\pi i \xi}
\Exp{\sum_{n>0}{1-t^{-n}\/n}p_n \xi^n}
\Exp{\sum_{n>0}(q^n-1){\partial\/\partial p_n} \xi^{-n}}
-{1\/t-1}.
\end{equation}
\noindent{\it Proof.\quad}
Since $q^{D_{x_i}} p_n = \((q^n-1)x_i^n+p_n\) q^{D_{x_i}}$,
we have
\begin{equation}
q^{D_{x_i}} =\,:\!\Exp{\sum_{n>0}(q^n-1)x_i^n{\partial\/\partial p_n}}\!:\,
= \oint{d\xi\/2\pi i \xi} \sum_{n\geq0} x_i^n \xi^n \cdot
\Exp{\sum_{n>0}(q^n-1){\partial\/\partial p_n} \xi^{-n}},
\end{equation}
here $:*:$ stands for the normal ordering such that
the differential operators ${\partial\/\partial p_n}$ are in the right.
It follows from eq.\ (III.2.9) and (III.2.10) in \cite{rM} that
\begin{equation}
\sum_i\prod_{j\neq i} {tx_i-x_j\/x_i-x_j}
\sum_{n\geq 0}x_i^n \xi^n
=
{t^N\/t-1}\Exp{\sum_{n>0}{1-t^{-n}\/n}p_n \xi^n}
-{1\/t-1}.
\end{equation}
This gives us the proposition. \hfill\fbox{}
Let
$\widetilde H_N(x_1,\cdots,x_N) \equiv t^{-N}\((t-1)H(x_1,\cdots,x_N)+1\)$,
then
\begin{equation}
\widetilde H_N(x_1,\cdots,x_N) \Pi(x,y) =
\widetilde H_M(y_1,\cdots,y_M) \Pi(x,y).
\end{equation}
With the self-adjointness of $H$ for the another scalar product,
we obtain the lemma 2 again.
|
\chapter{Field theory}
We can formulate scattering amplitudes in 4D field theory in a form close
to the one we use in string theory by means of the
Lehmann-Symanzik-Zimmermann reduction formula for $S$-matrix
elements~[\Ref{IZ}]:
$$
\eqalignno{& \langle \lambda_1, \ldots , \lambda_{N_{out}};
in \vert S \vert \lambda_{N_{out}+1}, \ldots,
\lambda_{N_{out}+N_{in}}, ;in
\rangle\ = \ {\rm disconnected \ terms}\ + &\nameali{LSZvo}\cr
&\qquad \prod^{N_{out}+N_{in}}_{j=1} \left( {i \over \sqrt{Z_j}}
\right)
\int \prod_{j=1}^{N_{out}+N_{in}}
\left( {{\rm d}}^4{x}_j \right) \langle 0 \vert {\rm T}
V_{\langle \lambda_1 \vert} (x_1) \ldots
V_{\vert \lambda_{N_{out}+N_{in}} \rangle} (x_{N_{\rm
out}+N_{in}}) \vert 0 \rangle \ . \cr}
$$
Here we have a Field Theory Vertex (FTV)
$V_{\vert \lambda \rangle} (x)$ corresponding
to the $1$-particle ket-state $\vert \lambda ; in \rangle$
where the label $\lambda$ incorporates the $4$-momentum $p$ as well as
other quantum numbers, and similarly we have a FTV
$V_{\langle \lambda \vert} (x)$ corresponding to the 1-particle bra-state
$\langle \lambda ; in \vert$.
Since by definition of hermitean conjugation
$\langle \lambda; in \vert = (\vert \lambda ; in \rangle )^{\dagger}$,
it is not surprising that $V_{\langle \lambda \vert}(x)$
is just the hermitean conjugate of $V_{\vert \lambda \rangle } (x)$,
$$ V_{\langle \lambda
\vert}(x) = \left( V_{\vert \lambda \rangle } (x) \right)^{\dagger} \ .
\nfr{qftinoutmap}
For example, for a particle described by a real scalar field $\phi$, the
$1$-particle states are specified by their momentum only and the Field
Theory Vertices are
$$\eqalignno{ V_{\vert p \rangle} (x) & = \ e^{i p \cdot x} \left( -
\square_x + m^2 \right) \phi(x) & \nameali{exone} \cr
V_{\langle p \vert} (x) & = \ e^{-i p \cdot x} \left( -
\square_x + m^2 \right) \phi(x) \ , \cr } $$
where in both cases $p^0 = + \sqrt{\vec{p}^{\, 2} + m^2}$ and
$\square = \eta^{\mu \nu} \del_{\mu} \del_{\nu}$
with $\eta = {\rm diag} (-1,1,1,1)$.
Another example is provided by an electron with momentum $p$ and helicity
$\eta$, where
$$\eqalignno{ V_{\vert e^-,p,\eta \rangle} (x) & = \ -
\overline{\psi} (x) \left( \buildchar{\slashchar\del_x}{\leftarrow}{}
- m \right) u(\vec{p},\eta) e^{i p \cdot x} & \nameali{extwo} \cr
V_{\langle e^-,p,\eta \vert} (x) & = \ -\overline{u} (\vec{p},\eta)
\left( -\slashchar\del_x - m \right) \psi(x) e^{-i p \cdot x} \ .
\cr} $$
Here $\overline{\psi} = \psi^{\dagger} (i \gamma^0) = \psi^{\dagger}
(-i \gamma_0)$ and $\{ \gamma^{\mu},\gamma^{\nu} \} = 2 \eta^{\mu \nu}$.
The spinor $u(\vec{p},\eta)$ of the incoming particle
with momentum $p$ and helicity $\eta$ satisfies
the Dirac equation $(i \slashchar{p} - m) u(\vec{p},\eta)=0$ and is
normalized according to
$$ u^{\dagger}(\vec{p},\eta) u(\vec{p},\eta') = 2 p^0 \delta_{\eta,\eta'} \ .
\nfr{ftspinornorm}
For particles of nonzero mass $m$ this normalization is equivalent to the
more standard one
$\overline{u}(\vec{p},\eta) u(\vec{p},\eta') = 2 m \delta_{\eta,\eta'}$,
but unlike the standard normalization condition it can also be used for
massless particles.
\chapter{String amplitudes}
In this paper we only consider 4D heterotic string models in a Minkowski
background.
We define the $T$-matrix element as the connected $S$-matrix element
with certain normalization factors removed
$$
\eqalignno{
& { \langle \lambda_1; \dots ; \lambda_{N_{out}}; in \vert S \vert
\lambda_{N_{out}+1}, \dots , \lambda_{N_{out} + N_{in}} ; in
\rangle_{\rm connected} \over
\prod_{i=1}^{N_{tot}} \left( \langle \lambda_i ; in \vert
\lambda_i ; in \rangle \right)^{1/2} }\ = & \nameali{Smatrix} \cr
&\qquad i (2\pi)^4 \delta^4 ( p_1+ \dots + p_{N_{out}} - p_{N_{out}+1} -
\dots - p_{N_{tot}} ) \ \prod_{i=1}^{N_{tot}}
(2 p_i^0 V)^{-1/2} \ \times \cr
&\qquad \ T ( \lambda_1; \dots ; \lambda_{N_{out}} \vert
\lambda_{N_{out}+1}; \dots ; \lambda_{N_{out} + N_{in}} ) \ , \cr }
$$
where $N_{tot} = N_{in} + N_{out}$ is the total number of
external states, $p_i$ is the momentum of the $i$'th string state, all
of them having $p_i^0 > 0$, and $V$ is the
usual volume-of-the-world factor. We also introduce the dimensionless
momentum $k_\mu \equiv \sqrt{{\alpha^\prime \over 2}} \, p_\mu$.
The Minkowski metric is $\eta = {\rm diag}(-1,1,1,1)$.
For heterotic superstrings in the Neveu-Schwarz Ramond formalism we have
various free conformal fields: The space-time coordinates $X^{\mu}$,
their chiral world-sheet superpartners $\psi^{\mu}$, the
reparametrization ghosts $b,c$ and $\bar{b},\bar{c}$, and the
superghosts $\beta,\gamma$. On top of this we have various internal
degrees of freedom described by a conformal field theory (CFT) with
left-moving (right-moving) central charge 22 (9). These may or may not
be free.
The $g$-loop contribution to the
$T$-matrix element is given by the Polyakov path
integral which is equivalent to the
following operator formula
$$
\eqalignno{ & T^g
(\lambda_1; \dots ; \lambda_{N_{out}} \vert
\lambda_{N_{out}+1};\dots;\lambda_{N_{out} + N_{in}} ) \ =
& \nameali{Tmatrix} \cr
&\qquad (-1)^{g-1} C_g \int \prod_{I=1}^{3g-3+N_{tot}}
\left( {\rm d}^2 m^I \right) \ \prod_{\mu=1}^{g}
\left(\sum_{{\bfmath\alpha}_{\mu},{\bfmath\beta}_{\mu}}
C^{{\bfmath\alpha}_{\mu}}_{{\bfmath\beta}_{\mu}} \right) \
\wew{\left| \prod_{I=1}^{3g-3+N_{tot}} (\eta_I \vert b)
\prod_{i=1}^{N_{tot}} c(z_i) \right|^2\ \times \cr
&\qquad \left(\prod_{A=1}^{2g-2+N_B+N_{FP}}
\Pi (w_A)\right) \, {\cal V}_{\langle \lambda_1 \vert } (z_1,\bar{z}_1)
\dots {\cal V}_{\vert \lambda_{N_{tot}} \rangle}
(z_{N_{tot}},\bar{z}_{N_{tot}})} \ . \cr}
$$
Here $C_g$ is a constant giving the proper normalization to the string
partition function (the $g$-loop vacuum amplitude). It will be given
explicitly in section 3, and (as we shall see) the sign
$(-1)^{g-1}$ ensures that $C_g$ is a positive number.
$m^I$ is a modular parameter,
$\eta_I$ is the corresponding Beltrami
differential, and our conventions for the overlap $(\eta_I \vert b)$
with the antighost
field $b$ are defined in detail in ref.~[\Ref{Kaj2}].
The integral is over one
fundamental domain of $N_{tot}$-punctured genus $g$ moduli space.
For each loop, labelled by $l=1,\ldots,g$, we have a summation over
sets of spin structures, collected in vectors ${\bfmath\alpha}_l$ and
${\bfmath\beta}_l$, and with a summation coefficient
$C^{{\bfmath\alpha}_l}_{{\bfmath\beta}_l}$.
By definition the correlator $\wew{\dots}$ includes the partition
function. At tree level, where the non-zero mode
partition function is equal to one,
the notation $\langle \dots \rangle$ is also used. At loop level we
choose the normalization for the partition function to be the one
obtained by applying the sewing procedure. This guarantees sensible
factorization properties in the corner of moduli space
where the world-sheet degenerates into individual tori connected by long
tubes and implies that the spin-structure summation coefficient is just
a product of one-loop summation coefficients as in eq.~\Tmatrix .
More details on our conventions for spin structures, partition functions and
operator fields in the explicit setting of a heterotic string model
built with free world-sheet fermions~[\Ref{KLT},\Ref{Anto},\Ref{Bluhm}]
can be found in Appendix A, see also refs.~[\Ref{ammedm},\Ref{mink}].
In analogy with field theory we have introduced a vertex operator
${\cal V}_{\vert \lambda \rangle} (z,\bar{z})$ for each ket string state
$\vert \lambda \rangle$ and similarly a vertex operator ${\cal
V}_{\langle \lambda \vert} (z,\bar{z})$ corresponding to each bra string
state $\langle \lambda \vert$.~\note{Since all the states we consider
are of the ``{\it in}'' variety, we drop the ``{\it in}'' label from now
on.} At the end of this section we will have more to say about the
meaning of these operators.
The ghost factors residing in the BRST invariant version of the
vertex operator, given by
$$ {\cal W}_{\vert \lambda \rangle} (z,\bar{z})\ =\
c(z) \overline{c}(\bar{z})
{\cal V}_{\vert \lambda \rangle} (z,\bar{z}) \qquad {\rm and} \qquad
{\cal W}_{\langle \lambda \vert} (z,\bar{z})\ =\
c(z) \overline{c}(\bar{z})
{\cal V}_{\langle \lambda \vert} (z,\bar{z}) \ ,
\nfr{wvvect}
have been factored out in eq.~\Tmatrix. We take all
space-time bosonic vertex operators to be in the
$q=-1$ superghost picture and all the space-time fermionic vertex
operators to be in the $q=-1/2$ superghost picture.
In an amplitude
involving $N_B$ space-time bosons and $2N_{FP}$
space-time fermions this implies that we have to insert $2g-2+N_B + N_{FP}$
PCOs $\Pi$ at
arbitrary points $w_A$ on the Riemann surface. In practical calculations
it can be convenient to insert one PCO at each of the vertex operators
describing the space-time bosons so as to change these into the $q=0$
picture. This leaves $2g-2+N_{FP}$ PCOs at
arbitrary points.
If we ``bosonize'' the superghosts in the usual way,
$\beta = \partial \xi e^{-\phi}$ and $\gamma = e^{+\phi}
\eta$, the PCO is given explicitly by
$$ \Pi = 2 c \partial \xi + 2 e^{\phi} T_F^{[X,\psi]} - {1 \over 2}
\partial (e^{2\phi} \eta b) - {1 \over 2} e^{2\phi} (\partial \eta) b \ ,
\nfr{PCO}
where we suppressed the superghost cocycle factor which ensures that
$e^{\phi}$ anti-commutes with all other fermionic operators on the
world-sheet, and
$$ T_F^{[X,\psi]} \ = \ - {i \over 2} \del X \cdot \psi + ({\rm internal
\ part }) \nfr{supcurr}
is the orbital part of the world-sheet
supercurrent (i.e. the part not involving ghosts and superghosts).
The ``internal part'' refers to the internal right-moving degrees
of freedom of the CFT with central charge $9$.
As stated in the introduction our aim in this paper is twofold:
First, since the $T$-matrix element as defined in eq.~\Smatrix\
corresponds to the connected $S$-matrix element obtained using states with
standard field theory normalization, we have to use vertex operators
with a definite normalization in eq.~\Tmatrix . So we need to know what
is the correct normalization of all vertex operators involved in the
theory; and we also need to determine the value of the overall
normalization constant $C_g$.
Second, we need to understand what is the {\sl exact} relation between the
vertex operators
${\cal W}_{\langle \lambda \vert}(z,\bar{z})$ and
${\cal W}_{\vert \lambda \rangle}(z,\bar{z})$.
By definition the operator ${\cal W}_{\vert
\lambda \rangle} (z=0)$, when acting on the conformal vacuum $\vert 0
\rangle$, creates the string state $\vert \lambda \rangle$, where (like
in section 1) $\lambda$ is a label incorporating the $4$-momentum $k$ (with
$k^0 > 0$), the helicity and the ``particle type'' (defined through the
values of various charges and family labels).
We may think of eq.~\Tmatrix\ as an indirect definition of what we mean
by ${\cal W}_{\langle \lambda \vert}$: It is the vertex operator we have
to use on the right-hand side of this equation in order to obtain the
$T$-matrix element involving the bra-state $\langle \lambda \vert =
(\vert \lambda \rangle)^{\dagger}$. Of course this definition is
somewhat circular, because we don't know how to compute the $T$-matrix
element until we have specified what are the vertex operators.
Indeed, as explained in the introduction,
the procedure we adopt is to carefully {\it derive}
what ${\cal W}_{\langle \lambda \vert}$
should be in order for the ``master formula'' \Tmatrix\ to reproduce
the correct amplitude for a propagating string to emit or
absorb a zero-momentum graviton.
Based on our experience from field theory, as outlined in section 1, we
might expect ${\cal W}_{\langle \lambda \vert}$ to be given by the
hermitean conjugate of ${\cal W}_{\vert \lambda \rangle}$. As we shall
see in section 4, this is not completely correct.
\chapter{Normalization of the vacuum amplitude}
In section 2 we already made use of the basic fact that
the problem of normalizing string amplitudes
can be separated into two independent problems: One, to fix the
normalization constant $C_g$ of the vacuum amplitude at genus $g$. The
other, to fix the normalization of each vertex operator in the theory.
It is factorization that leads to this simple result. For example, to
see that the normalization of the vertex operators cannot depend on
the topology of the world-sheet we can imagine
inserting a vertex operator on a sphere connected by a long tube to some
genus $g$ surface. It is clear that the vertex operator cannot
know about the distant handles. This is true even for vertex operators
describing space-time fermions, even though these involve spin fields
which are non-local operators on the world-sheet, because space-time
fermions always come in pairs and we may imagine isolating both of the
corresponding vertex operators (and the branch cut connecting them) on
a sphere far away from all handles.
Similarly, if we assume for the moment that the overall normalization of
the amplitude depends on the number $N$ of external states,~\note{In this
section only we drop the label {\it tot} on $N_{tot}$.}
as well as on the genus $g$, through some
coefficients $C_{g,N}$, we find by factorizing
the $N$-point $g$--loop amplitude into an $N+1$-point $g_1$--loop amplitude
times a $1$-point $g_2$--loop
amplitude times a propagator (where $g_1+g_2=g$), that
$$
C_{g_1+g_2,N}\ \propto\ C_{g_1,N+1}\ C_{g_2,1}
\efr
with a proportionality constant independent of $g_1$, $g_2$ and
$N$. Setting $g_1=0$ one gets
$$
C_{g,N}\ \propto\ C_{0,N+1}\ C_{g,1} \ ,
\efr
so that the dependence on $N$ can be studied at tree level.
Again by factorization, at tree level one gets
$$
C_{0,N_1+N_2}\ \propto\ C_{0,N_1+1}\ C_{0,N_2+1} \ ,
\efr
and if we put $N_2=2$ this implies that the ratio $C_{0,N+2} / C_{0,N+1}$
is independent of $N$
or, in other words, that $C_{0,N} \propto ({\cal M})^{N}$ for some constant
${\cal M}$. So we may write
$$
C_{g,N}\ =\ C_g\, ({\cal M})^{N} \ ,
\efr
and if we absorb a factor of ${\cal M}$
into the normalization of all vertex operators we are then left with an
overall normalization constant $C_g$ depending only on the genus.
To determine the value of $C_g$ we adopt the method proposed in
refs.~[\Ref{Kaj},\Ref{Marco}]:
To consider the elastic scattering of two gravitons in
the Regge regime of very high center-of-mass energy and small energy
transfer and impose that
the leading part of the $g$-loop amplitude assumes the universal form
needed for the eikonal resummation [\Ref{Veneziano}].
In order to get started we need the expression for the graviton vertex
operator including the proper normalization which was found in
refs.~[\Ref{Weinberg},\Ref{Kaj}]:
$${\cal V}^{(-1)}_{\vert {\rm grav} \rangle} (z,\bar{z}) = i
{\kappa \over \pi}
\bar{\epsilon} \cdot \bar{\del} X (\bar{z}) \epsilon \cdot \psi (z)
e^{-\phi(z)} e^{i k \cdot X(z,\bar{z})} \ , \nfr{gravvert}
where $k^2=0$ and we wrote the graviton polarization on the factorized
form $\bar{\epsilon} \otimes \epsilon$ with $\epsilon \cdot k =
\bar{\epsilon} \cdot k = 0$.
Our conventions for the operator fields can be found in Appendix A.
Like in eq.~\PCO\ we suppressed the cocycle factor which ensures that
the superghost
operator $e^{-\phi} = \delta(\gamma)$ anticommutes with all other
fermions on the world-sheet.
By picture changing \gravvert\ we arrive at
$$\eqalignno{ {\cal V}_{\vert {\rm grav} \rangle}^{(0)}
(z,\bar{z}) & = \ \lim_{w
\rightarrow z} \Pi(w) {\cal V}^{(-1)}_{\vert {\rm grav}
\rangle} (z,\bar{z}) \ &
\nameali{gravverttwo} \cr
& = \ {\kappa \over \pi} \bar{\epsilon} \cdot \bar{\del} X (\bar{z}) \left[
\epsilon \cdot \del X (z) - i k \cdot \psi (z) \epsilon \cdot \psi (z)
\right] e^{i k \cdot X (z,\bar{z})} \ . \cr } $$
The expressions for ${\cal V}^{(-1)}_{\langle {\rm grav} \vert}$ and
${\cal V}^{(0)}_{\langle {\rm grav} \vert}$ are identical to
eqs.~\gravvert\ and \gravverttwo , as long as the polarizations
$\epsilon,\bar{\epsilon}$ are taken to be real, and we ascribe to the
outgoing graviton a momentum with $k^0 < 0$.
The calculation of the four-graviton $g$-loop amplitude in the Regge
limit starting from eq.~\Tmatrix\
is different from the one in ref.~[\Ref{Kaj}] which was performed using
the manifestly world-sheet supersymmetric formulation of the heterotic
string.
In fact it is much harder, because even after changing the graviton
vertex operators into the $(0)$ picture there remains $2g-2$ PCOs at
arbitrary points. To obtain the universal form of the amplitude in the
pinching limit relevant for the Regge regime, where the world-sheet
degenerates into a ladder-like configuration consisting of two ``fast
legs'' connected by $g+1$ long tubes, one should insert $g-1$ PCOs on
each of the two ``fast legs''. (Other choices are of course possible but
will lead to the presence of total derivatives that make the leading
behaviour of the amplitude rather obscure.) Even subject to this
constraint there still remains $2g-2$ PCO insertion points, the
dependence on which only drops out at the very end of the calculation.
In the end we recover the standard result [\Ref{Kaj}] pertaining to $D=4$
space-time dimensions,
$$ C_g = \left( { 2\kappa^2 \over \alpha' } \right)^{g-1} \left( {1
\over 2\pi } \right)^{5g-3} (\alpha')^{-2} \nfr{overallnorm}
{\it and} the sign factor $(-1)^{g-1}$ explicitly displayed in
eq.~\Tmatrix. The origin of this sign is not too hard to understand. It is
needed to compensate the identical sign which
appears when we disentangle the anticommuting superghost factors
$e^{\phi}$ and
the orbital supercurrents $T_F^{[X,\psi]}$ in the product of the $2g-2$
PCOs
$$ \prod_{\alpha=1}^{2g-2} \left( e^{\phi(w_{\alpha})} T_F^{[X,\psi]}
(w_{\alpha}) \right) = (-1)^{g-1} \left(
\prod_{\alpha=1}^{2g-2} e^{\phi(w_{\alpha})} \right)
\left( \prod_{\alpha=1}^{2g-2} T_F^{[X,\psi]}
(w_{\alpha}) \right) \ . \efr
The other three terms present in the PCO \PCO\ do not contribute to the
leading behaviour of the amplitude in the Regge regime.
A comment about the spin structure summation coefficient
in eq.~\Tmatrix\ might be in order at
this point: We fix $C_g$ by considering the four-graviton $g$-loop
amplitude in the Regge regime. However, only the $2^g$ spin structures
responsible for graviton exchange contribute to the
leading, universal part of the amplitude.
How do we know that the
normalization we obtain is also correct for all the other spin
structures? The answer to this has already been given in section 2: The
requirement that the amplitude factorizes properly in the limit where
all loops are taken far apart implies that the spin structure summation
coefficient should be a product of one-loop summation coefficients.
These are in turn specified by the requirement that the one-loop
partition function should be modular invariant, once a (physically
sensible) choice of GSO projection has been
made~[\Ref{KLT},\Ref{Anto}].~\note{Strictly speaking modular invariance
of the one-loop partition function does not specify the summation
coefficient for those spin structures where one (or more) of the free
fermions on the world-sheet develop a zero mode, because these spin
structures give zero contribution to the partition function. In order to
check that no extra phase factors appear in these cases one may for
example consider the factorization of a two-loop vacuum amplitude into
one-loop tadpoles~[\Ref{Anto}]. We carried out this check explicitly in
the framework of Kawai-Lewellen-Tye~[\Ref{KLT}] heterotic string models.}
\chapter{The relation between ${\cal W}_{\vert \lambda \rangle}$ and
${\cal W}_{\langle \lambda \vert}$ }
We now consider in detail the connection between the vertex operators
describing incoming and outgoing string states.
What we are looking for is the map which, given the vertex operator
${\cal W}_{\vert \lambda \rangle}$
describing an incoming string state, gives us the vertex operator
${\cal W}_{\langle \lambda \vert}$
describing the same string state but outgoing. As we saw in section 1
this map is just given by hermitean conjugation in the framework of
quantum field theory. In string theory this cannot be the whole story,
because if the operator field ${\cal W}_{\vert \lambda \rangle}$
creates the ket-state $\vert \lambda \rangle$ in the usual sense,
$\vert \lambda \rangle =\lim_{\zeta,\bar{\zeta}\rightarrow 0}
{\cal W}_{\vert \lambda \rangle} (\zeta,\bar{\zeta}) \vert 0 \rangle $,
then by definition the
hermitean conjugate operator field creates the corresponding bra-state,
$\langle \lambda \vert =
\lim_{\zeta,\bar{\zeta}\rightarrow 0} \langle 0\vert
\left( {\cal W}_{\vert \lambda \rangle} (\zeta,\bar{\zeta})
\right)^\dagger$.
But in eq.~\Tmatrix\ both ${\cal V}_{\langle \lambda \vert}$ and
${\cal V}_{\vert \lambda \rangle}$ are vertex operators that create
ket states when acting on the conformal ket vacuum.
So we need to compose two-dimensional hermitean conjugation
with some other transformation which also maps a vertex operator
creating ket-states into a vertex operator creating bra-states.
This transformation should be a symmetry of any 2-dimensional conformal
field theory on the sphere. The obvious choice is the
{\sl BPZ conjugation\/} [\Ref{BPZ}] (see also [\Ref{Zwiebach}]).
Therefore we now quickly review our conventions on hermitean
conjugation and BPZ conjugation in conformal field theory.
After that we will propose a map from ${\cal W}_{\vert \lambda \rangle}$
to ${\cal W}_{\langle \lambda \vert}$
which is just an unknown phase
factor times the combination of
BPZ and hermitean conjugation. In the next section we will check
that our guess indeed gives the right map, and in the process the phase
factor will be determined.
\section{Two-dimensional hermitean conjugation}
In this section we review our conventions on hermitean conjugation, see
also refs.~[\Ref{Sonoda},\Ref{mink}]. We define the hermitean conjugate of all
elementary
operators in the conformal field theory by specifying the hermitean
conjugate of the corresponding oscillators, with the further
understanding that hermitean conjugation also complex conjugates all
complex numbers and inverts the order of the operators.
For example, if
$$\Phi_{\Delta} (z) = \sum_n \phi_n z^{-n-\Delta} \efr
is a primary chiral conformal field of conformal dimension $\Delta$, then the
hermitean conjugate of this field is
$$
\left(\Phi_{\Delta}(z)\right)^\dagger\ =\
\left({1\over z^*}\right)^{2\Delta}
\widehat\Phi_{\Delta}
({1\over z^*})\ ,
\nfr{hermconj}
where $z^*$ denotes the complex conjugate of $z$ (we think of $z$
and $\bar{z}$ as independent complex variables, so that $z^*$ and $\bar{z}$
need not be equal) and
$$ \widehat\Phi_{\Delta} (z) = \sum_n \phi_{-n}^{\dagger} z^{-n-\Delta}
\nfr{hermmode}
is a primary conformal field of the same dimension as $\Phi_{\Delta}$.
We say that a field $\Phi_{\Delta}$ is hermitean (anti-hermitean)
when $\widehat\Phi_{\Delta} = +\Phi_{\Delta} \ (-\Phi_{\Delta})$.
The hermiticity properties are made more complicated by the presence of
the reparametrization ghosts, because on the sphere the basic
nonvanishing correlator is
$ \langle \bar{c}_{-1} \bar{c}_0 \bar{c}_{1} c_{-1} c_0 c_1 \rangle $
where (since $c_n^{\dagger} = c_{-n}$)
the operator involved is explicitly anti-hermitean. Therefore
either one has to postulate an imaginary value for this correlator or
one has to relinquish the property $\langle M \vert A \vert N \rangle =
+ \langle N \vert A^{\dagger} \vert M \rangle^*$ of matrix elements
involving ghost degrees of freedom. We prefer the second option. We define
$$\langle \ \vert c_{-1} c_0 c_1 \vert^2 \ \rangle = \langle
\bar{c}_{-1} \bar{c}_0 \bar{c}_{1} c_{-1} c_0 c_1 \rangle = +1
\nfr{ghostone}
and this implies that
$$ \langle M \vert A \vert N \rangle = - \langle N \vert A^{\dagger}
\vert M \rangle^* \nfr{ghostherm}
in the presence of ghosts. As a special case of this
$$ \langle M \vert c_0 \bar{c}_0 A \vert N \rangle = \langle N \vert c_0
\bar{c}_0 A^{\dagger} \vert M \rangle^* \nfr{ghosttwo}
for any operator $A$ not involving the modes $b_0$ or $\bar{b}_0$.
A list of hermiticity properties for the fields relevant in
four-dimensional heterotic string models constructed using free fermions
can be found in Appendix B.
\section{BPZ invariance in conformal field theories}
Consider a conformal field theory on the cylinder.
Introduce complex coordinates $z=\exp\{ i (\sigma+\tau)\}$ and
$\bar{z} = \exp\{i(-\sigma+\tau)\}$ and rotate
to Euclidean time $\tau \rightarrow -i \tau$.
Changing sign on $\tau$ and $\sigma$ simultaneously gives rise to the
Belavin-Polyakov-Zamolodchikov (BPZ) transformation $z \rightarrow 1/z$
[\Ref{BPZ},\Ref{Zwiebach}]. This transformation
defines a globally holomorphic diffeomorfism on the sphere.
At the level of the operator
fields, the transformation changes the coordinate system from $(z)$ to
$(w)$ where $w=1/z$:
$$ \Phi (z=\zeta) \bpzarrow \Phi (w=\zeta) \ .
\nfr{bpz}
For a primary conformal field of dimension $\Delta$
$$ \Phi_{\Delta} (w=\zeta) =
e^{-i \epsilon \pi \Delta} \left( {1 \over \zeta} \right)^{2\Delta}
\Phi_{\Delta} \left( z= {1 \over \zeta} \right) \ , \nfr{bpzprimary}
where for non-integer conformal dimensions we have to
choose a specific phase for $-1$,~\note{In
their original paper~[\Ref{BPZ}], Belavin, Polyakov and
Zamolodchikov avoided this problem by considering instead the conformal
transformation $z \rightarrow -1/z$, but we prefer to consider
$z \rightarrow 1/z$, in accordance with most subsequent authors.}
parametrized by an odd integer
$\epsilon$, when forming the transformation factors
$$ {dz \over dw} = e^{-i \epsilon \pi} {1 \over w^2} \qquad {\rm and}
\qquad
{dw \over dz} = e^{+i \epsilon \pi} {1 \over z^2} \ . \nfr{transfac}
The BPZ transformation does not reverse the order of operators and it
leaves all complex numbers unchanged. It cannot itself be generated by
any operator acting on ket states.
Instead it defines a map from ket-states to bra-states as follows:
$$ \vert \Phi \rangle \ \equiv \ \lim_{\zeta \rightarrow 0}
\Phi (z=\zeta) \vert 0 \rangle \ \ \bpzarrow \ \
\langle \Phi^{\rm BPZ} \vert \ \equiv \ \lim_{\zeta \rightarrow 0}
\langle 0 \vert \Phi (w=\zeta) \ . \nfr{bpzstate}
The label ``BPZ'' on the state $\langle \Phi^{\rm BPZ} \vert$ is
necessary in order to avoid confusion with the bra state
$\langle \Phi \vert \equiv
\lim_{\zeta \rightarrow 0} \langle 0 \vert \left( \Phi
(z=\zeta) \right)^{\dagger}$
defined by hermitean conjugation, because this will in general
differ from $\langle \Phi^{\rm BPZ} \vert$. (Another possibility,
preferred by many authors, is to take BPZ conjugation as the defining
map from ket to bra and introduce instead a label $\langle \Phi^{\rm
h.c.} \vert$ on the state defined by hermitean conjugation.)
\section{Composing BPZ and hermitean conjugation}
The composition of BPZ and hermitean conjugation gives a map from
ket to ket
$$
\vert\Phi \rangle\
\longrightarrow\
\left(\langle\Phi^{\rm BPZ}\vert\right)^\dagger\ =\
\vert \Phi^{\rm BPZ}\rangle\ \efr
which acts on the primary conformal fields as follows
$$ \eqalignno{
\Phi_{\Delta,\bar{\Delta}}(z=\zeta,\bar{z}=\bar{\zeta})\ \longrightarrow\
& \left( \Phi_{\Delta,\bar{\Delta}} (w=\zeta,\bar{w}=\bar{\zeta})
\right)^{\dagger} & \nameali{hermandbpz} \cr
& \ = \
e^{i\epsilon\pi(\Delta-\bar{\Delta})}
\widehat\Phi_{\Delta,\bar{\Delta}}
(z=\zeta^*,\bar{z}=\bar{\zeta}^*) \ . \cr} $$
Notice that for fields with non-integer value of
$\Delta-\bar{\Delta}$,
BPZ and hermitean conjugation do not commute.
However, this is not a
problem for vertex operators describing BRST-invariant on-shell string
states, which satisfy $\Delta=\bar{\Delta}=0$.
The transformation \hermandbpz\ is our educated guess for the map taking
${\cal W}_{\vert \lambda \rangle}$ into
${\cal W}_{\langle \lambda \vert}$, only we will allow the possibility
that some phase factor $\chi$ may appear.
In other words, our ansatz is that if some incoming string
state with definite quantum numbers is created, in the superghost charge
$q$ picture, by the vertex operator
${\cal W}_{\vert\lambda\rangle}^{(q)}$,
$$
\vert \lambda \rangle\ =\ \lim_{z,\bar{z} \rightarrow 0}
{\cal W}_{\vert\lambda\rangle}^{(q)} (z,\bar{z}) \vert 0 \rangle \ ,
\efr
then the vertex operator we have to use in the ``master formula''
\Tmatrix\ to obtain the $T$-matrix element involving the outgoing state
$\langle \lambda \vert$ is given by
$$ {\cal W}_{\langle \lambda \vert}^{(q)} (z=\zeta,\bar{z}=\bar{\zeta})
\ \equiv\ \chi_q \left( {\cal W}_{\vert \lambda \rangle}^{(q)}
(w=\zeta^*,\bar{w} =\bar{\zeta}^*) \right)^{\dagger} \ .
\nfr{guesstwo}
As was emphasized at the beginning of section 4, the operator ${\cal
W}_{\langle \lambda \vert}^{(q)}$, like any vertex operator, creates a
state by acting on the ket vacuum. From the definitions \guesstwo\ and
\bpzstate\ we find this state to be
$$ \lim_{\zeta,\bar{\zeta} \rightarrow 0} {\cal
W}_{\langle \lambda \vert}^{(q)} (z=\zeta,\bar{z}=\bar{\zeta})
\vert 0 \rangle \ = \
\chi_q \vert \lambda^{\rm BPZ} \rangle \ .
\efr
In other words, we obtain the $T$-matrix element involving the bra-state
$\langle \lambda \vert$ by inserting into the Polyakov path integral an
operator creating the state $\chi_q \vert \lambda^{\rm BPZ} \rangle$.
Notice that whereas the state $\vert \lambda \rangle$ always has
$k^0 > 0$, the state $\vert \lambda^{\rm BPZ} \rangle$ has $k^0 < 0$.
Since the combination of BPZ and hermitean conjugation maps
$$ L_0 \rightarrow L_0 \qquad {\rm and} \qquad
Q_{BRST} \rightarrow - Q_{BRST} \ , \efr
and since BPZ conjugation is a world-sheet symmetry on the sphere, it
follows that if the state $\vert \lambda \rangle$ is
a physical on-shell state,
$L_0 \vert \lambda \rangle = Q_{BRST} \vert \lambda \rangle = 0$,
then so is the state $\chi_q \vert \lambda^{\rm BPZ} \rangle$,
regardless of what value we choose for the phase $\chi_q$. It
is less clear that the map \guesstwo\ is also
consistent with the GSO projection, i.e. that $ \vert \lambda \rangle$
satisfies the GSO projection conditions if and only if
$\vert \lambda^{\rm BPZ} \rangle$ does,
because the two states will in general reside in
different sectors of the string theory. An explicit proof in the
framework of a Kawai-Lewellen-Tye (KLT) type
heterotic string model is given in Appendix C.
If we restrict ourselves to BRST invariant on-shell string states, both
${\cal W}_{\vert \lambda \rangle}$ and $\widehat{\cal W}_{\vert
\lambda \rangle}$ are primary conformal fields of dimension zero, and
eq.~\guesstwo\ becomes
$$ {\cal W}^{(q)}_{\langle \lambda \vert} (\zeta,\bar{\zeta}) = \chi_q
\widehat{\cal W}^{(q)}_{\vert \lambda \rangle} (\zeta,\bar{\zeta}) \ .
\nfr{guessthree}
We will now proceed to verify our ansatz \guessthree\
by considering the amplitude
for the string state $\vert \lambda \rangle$ to emit (absorb) a very
soft graviton. We will find that the phase factor $\chi_q$, as
anticipated by our notation, depends only
on the choice of picture. In particular, if we restrict ourselves to the
pictures $q=-1$ and $q=-1/2$, the phase factor $\chi_q$ depends only on
whether the string state is a space-time boson or a space-time
fermion. At the same time we will be able to determine the correct
overall normalization of the vertex operators to be used in the formula
\Tmatrix\ for the $T$-matrix element.
\chapter{Normalization of vertex operators}
In this section we consider the computation of the tree-amplitude for
some given on-shell string state to absorb or emit a very soft graviton.
We perform the
analysis for a generic four dimensional heterotic string theory
where the graviton vertex operator has the form of eq.~\gravvert, but
the argument can be readily applied to
other string models.
We first discuss the case of space-time bosonic states and then
the case of the space-time fermionic states.
\section{Normalization of space-time bosonic vertex operators}
We first recall what is the situation in field theory.
Consider a basis of propagating bosonic particle states with momentum
$p$, labelled by an index $N$, in terms of which the
propagator assumes the diagonal form $P_{MN}/(p^2+ m_N^2)$ where
$P_{MN} = + \delta_{M,N}$ for physical states and
$P_{MN} = -\delta_{M,N}$ for possible
negative norm states. For example, for a photon with space-time vector
index $M=\mu$ we have $P_{MN} = \eta_{\mu \nu}$.
The tree-level $T$-matrix element for such a
particle to emit (absorb) a graviton contains a universal term which, in
the limit where the graviton momentum is zero, assumes the form
$$
\eqalignno{-2\kappa \, \epsilon \cdot p \ \bar\epsilon \cdot p
\ P_{MN}
&=\ - 4 {\kappa\over \alpha^\prime} \, \epsilon \cdot k \
\bar\epsilon \cdot k \ P_{MN} &\nameali{rightres}\cr
&=\ - C_0 \left(\kappa\over\pi\right)^3
\, \epsilon \cdot k \ \bar\epsilon \cdot k \ P_{MN} \ , \cr}
$$
where we wrote the graviton polarization on the factorized form
$\epsilon\otimes\bar\epsilon$ and $C_0$ is the overall
normalization constant for string tree amplitudes, given by
eq.~\overallnorm. The behaviour \rightres\ describes the canonical
coupling of gravity to the $p_{\mu} p_{\nu}$-part of the energy-momentum
tensor of the propagating particle.
The sign of the amplitude \rightres\ obviously depends on the sign
convention for the graviton field $h_{\mu \nu}$.
Eq.~\rightres\ corresponds to the
expansion
$$ g_{\mu \nu} = \eta_{\mu \nu} - 2 \kappa \left( h_{\mu \nu} + \lambda
\eta_{\mu \nu} h^{\sigma}_{\ \sigma} \right) + {\cal O} (h^2)
\nfr{gravfield}
regardless of the coefficient $\lambda$ chosen for the trace term.
The sign chosen for the graviton vertex operator \gravvert\
is in agreement with
this convention, as one may check by computing the 3-graviton tree
amplitude from eqs.~\gravvert\ and \Tmatrix\ and comparing with
eq.~\rightres\ in the case where the state $\vert M \rangle$ is itself
a graviton.
Consider now computing the universal part \rightres\ of the graviton
absorption amplitude at genus zero in string theory. We consider a
complete set of space-time bosonic string states $\vert N,k \rangle$,
labelled by $N$, built from the superghost vacuum $\vert q=-1\rangle$,
satisfying $b_0 = \bar{b}_0 = 0$ and having definite momentum $k$.
We may think of $N$ as specifying physical quantities such as
helicity, charges and family labels.
The $T$-matrix element for the process ``$N + {\rm graviton \ } \rightarrow
M$'' is given by
$$
T^0(M,k \vert {\rm graviton}; N,k)\ =\ -C_0
\vev{ {\cal W}_{\langle M,k \vert}^{(-1)}(z_1,\bar{z}_1)
{\cal W}_{\vert {\rm grav} \rangle}^{(0)}(z,\bar{z})
{\cal W}_{\vert N,k \rangle}^{(-1)}(z_2,\bar{z}_2) } \ ,
\nfr{mgravn}
where we have to use the graviton vertex operator in the superghost
charge $(0)$ picture, given by eq.~\gravverttwo ,
and the states $\vert N,k \rangle$ and $\langle
M,k
\vert$ are now assumed to be physical, so that ${\cal W}^{(-1)}_{\vert
N,k
\rangle}$ and ${\cal W}_{\langle M,k \vert}^{(-1)}$ are primary conformal
fields of dimension zero.
By projective invariance on the sphere we can fix
$z_1=\infty$, $z=1$ and $z_2=0$; and since the
${\cal W}_{\langle M,k \vert}^{(-1)}$
vertex operator is assumed to have conformal dimension zero
we can evaluate it in the coordinate system $(w)$, where
$w=1/z$, without introducing any transformation factor.
In so doing we just undo the BPZ transformation in the definition
eq.~\guesstwo\ of the operator ${\cal W}_{\langle M,k \vert}^{(-1)}$
and obtain
$$ \langle 0 \vert {\cal W}^{(-1)}_{\langle M,k \vert} (w=\bar{w}=0) \ = \
\chi_{-1} \ \langle 0 \vert \left( {\cal W}_{\vert M,k
\rangle}^{(-1)} (z=\bar{z}=0) \right)^{\dagger} \ = \ \chi_{-1} \
\langle M,k \vert \ .
\nfr{fivefour}
Accordingly eq.~\mgravn\ becomes
$$
T^0 (M,k \vert {\rm graviton}; N,k) \ = \ -\chi_{-1} \ C_0
\left({\kappa\over\pi}\right) \vev{ M,k \vert c(1) \bar{c}(1)
\bar\epsilon\cdot\bar\del X (1)
\, \epsilon\cdot\del X(1) \vert N,k } \ . \nfr{mgravnthree}
Here we may expand the fields $c$, $\bar{c}$, $\partial X$ and
$\bar{\partial} X$ in oscillators. Only modes with $L_0 = \bar{L}_0=0$ can
contribute to the ``universal'' part \rightres\ of the amplitude. This
is because this part of the amplitude, like that of a freely propagating
string state, conserves $L_0(X^{\mu})$, $\bar{L}_0 (X^{\mu})$, $L_0(b,c)$
and $\bar{L}_0 (\bar{b},\bar{c})$. We may imagine the basis
$\vert N,k \rangle$ of string states to diagonalize all these operators.
Then for $n \neq 0$ we may write e.g.
$$ \alpha_n^{\mu} = - {1 \over n} \left[ L_0 (X^{\mu}) ,
\alpha_n^{\mu} \right]
\efr
and this vanishes between the states $\langle M,k \vert$ and $\vert N,k
\rangle$ since by assumption they have the same value of $L_0 (X^{\mu})$.
We are thus left with
$$ T^0 ( M,k \vert {\rm graviton}; N,k ) = -
\chi_{-1} \ C_0 \left({\kappa \over
\pi}\right) \ \epsilon \cdot k \ \bar{\epsilon} \cdot k \
\vev{ M,k \vert \bar{c}_0 c_0 \vert N,k } + \ldots \ , \nfr{mgravntwo}
where ``$+ \ldots$'' denotes possible other terms in the amplitude with a
different kinematical structure than the universal part \rightres .
By eq.~\ghosttwo\ the matrix $\vev{ M,k \vert \bar{c}_0 c_0 \vert N,k }$ is
manifestly hermitean and by an appropriate choice of basis it may be
diagonalized such that
$$ \vev{ M,k \vert \bar{c}_0 c_0 \vert N,k } = \left(
{\cal N}_{\vert M,k\rangle}^{\rm bos}\right)^*
{\cal N}_{\vert N,k \rangle}^{\rm bos} \ P_{MN} \ ,
\nfr{statenorm}
where either $P_{MN} = 0$ (so that the state does not propagate) or
$|P_{MN}| = \delta_{M,N}$. Our conventions \ghostone\ imply that
$P_{MN}=+\delta_{M,N}$ for all physical external states but
$-\delta_{M,N}$ for
negative norm states (such as the ``timelike'' photon). The factor
${\cal N}_{\vert N,k \rangle}^{\rm bos}$
specifies the overall normalization of the
state $\vert N,k \rangle$. By inserting
eq.~\statenorm\ into eq.~\mgravntwo\ we obtain finally the correct
result \rightres\ {\it if} we take the phase factor introduced in
eq.~\guesstwo\ to be $\chi_{-1}=1$ and choose the normalization constant
to be the same for all states,
${\cal N}_{\vert M,k\rangle}^{\rm bos}={\cal N}_{\vert N,k\rangle}^{\rm
bos}$, given by
$$\left| {\cal N}_{\vert N,k \rangle}^{\rm bos} \right|
= {\kappa \over \pi} \ . \efr
In summary,
$$ {\cal W}^{(-1)}_{\langle N,k \vert} (z,\bar{z}) = + \widehat{\cal
W}^{(-1)}_{\vert N,k
\rangle} (z,\bar{z}) \qquad {\rm for \ physical \ spacetime \ bosons} \ ,
\nfr{inoutboson}
and the proper normalization of the state $\vert N,k \rangle$ is given by
$$ \langle M,k \vert \bar{c}_0 c_0 \vert N,k \rangle = \left( {\kappa \over
\pi } \right)^2 P_{MN} \ . \nfr{statenormtwo}
Since by definition
$ \vert N,k \rangle = \lim_{\zeta,\bar{\zeta} \rightarrow 0} {\cal
W}_{\vert N,k \rangle} (\zeta,\bar{\zeta}) \vert 0 \rangle $,
eq.~\statenormtwo\ specifies the normalization of the vertex operator up
to a complex phase factor. If the vertex operator
${\cal W}_{\vert N,k \rangle}^{(-1)}$ is complex, i.e. not proportional to
$\widehat{\cal W}_{\vert N,-k \rangle}^{(-1)}$,~\note{Notice that if
${\cal W}_{\vert N,k \rangle}^{(-1)}$ is proportional to $\exp(ik\cdot X)$
then $\widehat{\cal W}_{\vert N,k \rangle}^{(-1)}$ is proportional to
$\exp(-ik\cdot X)$.} there is probably no
fundamental reason to prefer any specific value of the overall complex
phase factor, just as in field theory the phase of a complex field is
an unphysical degree of freedom. If, on the other hand,
${\cal W}_{\vert N,k \rangle}^{(-1)}$ {\it is} proportional to
$\widehat{\cal W}_{\vert N,-k \rangle}^{(-1)}$ it becomes natural to
impose a reality condition, which we can take to be
$$ {\cal
W}_{\vert N,k \rangle}^{(-1)} = + \widehat{\cal
W}_{\vert N,-k \rangle}^{(-1)} \qquad {\rm or} \qquad
{\cal V}_{\vert N,k \rangle}^{(-1)} = - \widehat{\cal
V}_{\vert N,-k \rangle}^{(-1)} \efr
in agreement with the choice made for the graviton vertex operator
\gravvert . This implies that ${\cal W}_{\langle N,k \vert}^{(-1)} =
{\cal W}_{\vert N,-k \rangle}^{(-1)}$. Even in this case there remains a
choice a sign for the vertex operator. This is completely dependent on
convention, just like the sign of the graviton field in the expansion
\gravfield .
\section{Normalization of space-time fermionic vertex operators}
We now consider the case of space-time fermions. The field theory
description is now more complicated than in the case of space-time
bosons, since the graviton field should be described in terms of the
vierbein, $e^{\mu}_m$. The canonical coupling to gravity of a Dirac
fermion, labelled by an index $N$, is given by the action
$$ \int {\rm d}^4 x\ e \ \overline{\psi}_M \left\{ \gamma^m e^{\mu}_m
\partial_{\mu} + m \right\} \psi_N \ P^{MN} \ , \efr
where we ignore the spin-connection terms which all involve derivatives
of the vierbein and thus give rise to terms in the
fermion-fermion-graviton amplitude proportional to the graviton
momentum. When expanding $e^{\mu}_m$ around the flat background we can
ignore the deviation of $e=\det \{ e^{\mu}_m \}$ from unity since this
gives rise only to terms proportional to the trace of the graviton
field.
One obtains the following expression, analogous to eq.~\rightres\ for
the universal part of the fermion-fermion-graviton $T$-matrix element at
tree level:
$$ -i \kappa \overline{u}(\vec{p},\eta)
\gamma^{\nu} p^{\mu} u (\vec{p},\eta) \epsilon_{\nu} \bar{\epsilon}_{\mu}
\ P_{MN} \ ,
\nfr{fermrightres}
where, by virtue of the Gordon identity
$$\overline{u}(\vec{p},\eta) \gamma^{\nu} u(\vec{p},\eta') = -2i p^{\nu}
\delta_{\eta,\eta'} \ ,
\nfr{qftgordon}
we recover the bosonic result \rightres , as dictated by the
principle of equivalence.
In the string theory analysis we again consider a complete set of
states $\vert N,k \rangle$, labelled by $N$,
now built from the superghost vacuum $\vert q=-1/2 \rangle$,
again satisfying $b_0 = \bar{b}_0 = 0$ and having a definite momentum
$k$.
We may now proceed exactly as in section 5.1, only now we have to use
the superghost charge $(-1)$ version of the graviton vertex operator,
given by eq.~\gravvert.
In the limit of vanishing graviton momentum we obtain
$$
\eqalignno{ T^0(M,k \vert {\rm graviton} ; N,k)\ &= \
-C_0 \ \vev{ {\cal W}_{\langle M,k \vert}^{(-1/2)}(z_1,\bar{z}_1)
{\cal W}_{\vert {\rm grav} \rangle}^{(-1)}(z,\bar{z})
{\cal W}_{\vert N,k \rangle}^{(-1/2)}(z_2,\bar{z}_2) } \cr
& = \ - \chi_{-1/2} \ C_0 \ \langle M,k \vert
{\cal W}_{\vert {\rm grav} \rangle}^{(-1)}(1) \vert N,k \rangle \ .
&\nameali{vmandvnone} \cr } $$
As in the bosonic case only zero-mode operators contribute to the part
of the amplitude in which we are interested, so that
$$ T^0(M,k \vert {\rm graviton} ; N,k) \ = \
\chi_{-1/2} \ C_0 {\kappa \over \pi } \ \bar{\epsilon} \cdot k \
\epsilon_{\nu} \langle M,k \vert
\bar{c}_0 c_0 \psi_0^{\nu} \delta(\gamma_0)
\vert N,k \rangle + \ldots \ .
\nfr{vmandvn}
Here we may recognize the form \fermrightres\ of the result obtained in
field theory, since the zero mode $\psi_0^{\nu}$ of the operator field
$\psi^{\nu}$ furnishes a representation of the Clifford algebra, and so
is completely analogous to the gamma matrix $\gamma^{\nu}$ appearing in
the expression \fermrightres .
The matrix $\langle M,k \vert \bar{c}_0 c_0 \psi_0^{\nu} \delta(\gamma_0)
\vert N,k \rangle$ transforms as a space-time vector and therefore has to
be proportional to the momentum $k^{\nu}$. Since $\psi_0^{\nu}$ and
$\delta(\gamma_0)$ anti-commute it is manifestly
anti-hermitean (q.v. eq.~\ghosttwo )
and by choosing an appropriate basis it can be
diagonalized such that
$$\langle M,k \vert \bar{c}_0 c_0 \psi_0^{\nu} \delta(\gamma_0)
\vert N,k \rangle = i Y k^{\nu} \left(
{\cal N}_{\vert M,k \rangle}^{\rm ferm} \right)^*
{\cal N}_{\vert N,k \rangle}^{\rm ferm} \ P_{MN} \ .
\nfr{fermnorm}
In section 7 we will explicitly derive this formula in the context of a
KLT heterotic string model. It is quite analogous to eq.~\statenorm .
The factor of $i$ reflects the fact that the matrix on the left-hand
side is anti-hermitean and (as we shall see in section 7)
the constant factor $Y=\pm 1$ depends on the conventions chosen for the
spin fields.
Finally, $P_{MN} = + \delta_{M,N}$ for physical states, as always.
Like in the bosonic case the factor
${\cal N}_{\vert N,k \rangle}^{\rm ferm}$ specifies the normalization of
the string state $\vert N,k \rangle$.
If we insert eq.~\fermnorm\ into eq.~\vmandvn\ we finally obtain
$$\eqalignno{
& T^0 ( M,k \vert {\rm graviton}; N,k ) & \nameali{fivetwenty} \cr
& \qquad \qquad = \
Y \ i \ \chi_{-1/2} \ C_0 \left(
{\cal N}_{\vert M,k \rangle}^{\rm ferm} \right)^*
{\cal N}_{\vert N,k \rangle}^{\rm ferm}
\left( \kappa \over \pi \right)
\bar{\epsilon} \cdot k \ \epsilon \cdot k \ P_{MN} + \ldots \ , \cr} $$
which reproduces the right result \rightres\ assuming we choose
$$ \chi_{-1/2} = i Y \qquad {\rm for \ spacetime \ fermions}
\nfr{ysign}
and fix the normalization of the states in the same universal way as for
the bosons,
${\cal N}_{\vert M,k \rangle}^{\rm ferm}={\cal N}_{\vert N,k \rangle}^{\rm
ferm}$, and
$$ \left| {\cal N}_{\vert N,k \rangle}^{\rm
ferm} \right| = \left| {\cal N}_{\vert N,k \rangle}^{\rm
bos} \right| = {\kappa \over \pi} \ . \nfr{uninorm}
In summary
$$
{\cal W}_{\langle N,k \vert}^{(-1/2)} (z,\bar{z})\ =\ (i Y)\,
\widehat{\cal W}_{\vert N,k \rangle}^{(-1/2)}(z,\bar{z}) \qquad
{\rm for \ physical \ spacetime \ fermions,}
\nfr{fermap}
and the proper normalization of the string state is given by
$$\langle M,k \vert \bar{c}_0 c_0 \psi_0^{\nu} \delta(\gamma_0)
\vert N,k \rangle = i Y k^{\nu} \left( {\kappa \over \pi } \right)^2
\ P_{MN} \ .
\nfr{fermnormtwo}
Since the PCO \PCO\ is an anti-hermitean operator which satisfies Bose
statistics, eqs.~\inoutboson\ and \fermap\ can be generalized to the
superghost charge $q$ picture as follows
$$ {\cal W}_{\langle N,k \vert}^{(q)} (z,\bar{z}) \ = \ \chi_{q} \
\widehat{\cal W}_{\vert N,k \rangle}^{(q)} (z,\bar{z}) \ = \
(-1)^{q+1} \ \widehat{\cal W}_{\vert N,k \rangle}^{(q)} (z,\bar{z})
\ . \nfr{inoutq}
For pictures of half-integer $q$ (i.e. pictures describing space-time
fermions) the phase factor $(-1)^{q+1}$ involves a choice of sign, which
is parametrized by $Y$ according to eq.~\ysign , i.e. $(-1)^{1/2} = iY$.
\chapter{Space-Time hermiticity}
An important check on the correctness of our expressions
\inoutboson\ and \fermap\ for ${\cal W}_{\langle N,k \vert}^{(q)}$
is provided by the requirement that
the $T$-matrix element obtained from eq.~\Tmatrix\ has the right
hermiticity properties.
Unitarity requires that the tree-level $T$-matrix element is real except
when the momentum flowing in some intermediate channel happens to be on
the mass-shell corresponding to some physical state in the theory.
In field theory the imaginary part appears as a result of the
$i\epsilon$-prescription present in the propagator that happens to be
on-shell. In string theory it appears as a result of some divergency in
the integral over the Koba-Nielsen (KN) variables that has to be treated in a
way consistent with the $i\epsilon$-prescription in field
theory~[\Ref{Hoker},\Ref{Berera},\Ref{Weisberger}].
What we can rather easily show is that as long as the integrals over the
KN variables are convergent the expressions
\inoutboson\ and \fermap\ lead to a hermitean $T$-matrix at tree level.
At genus zero the formula \Tmatrix\ can be rewritten as
$$
\eqalignno{ & T^0
(\lambda_1; \dots ; \lambda_{N_{out}} \vert
\lambda_{N_{out}+1};\dots;\lambda_{N_{out} + N_{in}} ) \ =
& \nameali{Ttree} \cr
&\qquad - C_0 \ \int \left( \prod_{i=4}^{N_{tot}}
{\rm d}^2 z_i \right) \
\vev{ \bar{c}(\bar{z}_1) \bar{c}(\bar{z}_2) \bar{c}(\bar{z}_3)
c(z_1) c(z_2) c(z_3)\ \times \cr
&\qquad\quad \left(\prod_{A=1}^{N_B+N_{FP}-2}
\Pi (w_A)\right) \, {\cal V}_{\langle \lambda_1 \vert} (z_1,\bar{z}_1)
\dots {\cal V}_{\vert \lambda_{N_{\rm tot}} \rangle}
(z_{N_{\rm tot}},\bar{z}_{N_{\rm tot}})} \ . \cr}
$$
The $T$-matrix is hermitean if and only if the quantity \Ttree\ equals
$$
\eqalignno{ & \left[ T^0 ( \lambda_{N_{out} + N_{in}}; \ldots ;
\lambda_{N_{out}+1} \vert \lambda_{N_{out}} ; \ldots ;
\lambda_1 ) \right]^*\ = & \nameali{Ttreeherm} \cr
&\qquad + C_0 \ \int \left( \prod_{i=4}^{N_{\rm tot}}
{\rm d}^2 z_i^* \right) \
\vev{ \left( {\cal V}_{\vert \lambda_1 \rangle} (z_1,\bar{z}_1)
\right)^{\dagger} \ldots \left(
{\cal V}_{\langle \lambda_{N_{\rm tot}} \vert}
(z_{N_{\rm tot}},\bar{z}_{N_{\rm tot}}) \right)^{\dagger}\ \times \cr
&\qquad\ \left( \Pi (w_{N_B + N_{FP}-2}) \right)^{\dagger} \ldots
\left( \Pi(w_1) \right)^{\dagger} \ \left( c(z_3) \right)^{\dagger}
\dots \left( \bar{c}(\bar{z}_1) \right)^{\dagger} } \ , \cr }
$$
where we used eq.~\ghostherm.
In terms of the vertex operators ${\cal V}$ (where the $c\bar{c}$ factor
present in ${\cal W}$ has been removed, q.v. eq.~\wvvect) the relations
\inoutboson\ and \fermap\ acquire an extra
minus sign (because $c\bar{c}$ is an anti-hermitean operator):
$$\eqalignno{ {\cal V}^{(-1)}_{\langle \lambda \vert} (z,\bar{z}) & = \
- \widehat{\cal V}_{\vert \lambda \rangle}^{(-1)} (z,\bar{z}) &
\nameali{inoutv} \cr
{\cal V}^{(-1/2)}_{\langle \lambda \vert} (z,\bar{z}) & = \
- i Y \ \widehat{\cal V}_{\vert \lambda \rangle}^{(-1/2)} (z,\bar{z})
\ , \cr } $$
which, by taking the hermitean conjugate, leads to the inverse relations
$$\eqalignno{ {\cal V}^{(-1)}_{\vert \lambda \rangle} (z,\bar{z}) & = \
- \widehat{\cal V}_{\langle \lambda \vert}^{(-1)} (z,\bar{z}) &
\nameali{outinv} \cr
{\cal V}^{(-1/2)}_{\vert \lambda \rangle} (z,\bar{z}) & = \
- i Y \ \widehat{\cal V}_{\langle \lambda \vert}^{(-1/2)} (z,\bar{z})
\ . \cr } $$
Since the operators ${\cal V}_{\vert \lambda \rangle}$ and
${\cal V}_{\langle \lambda \vert}$ have conformal dimensions
$\Delta=\bar{\Delta} = 1$ we find for $i=1,\ldots,N_{out}$:
$$\eqalignno{ \left( {\cal V}_{\vert \lambda_i \rangle}
(z_i,\bar{z}_i) \right)^{\dagger} \ &= \ \left( {1 \over z_i^*} {1 \over
\bar{z}_i^*} \right)^2 \widehat{\cal V}_{\vert \lambda_i \rangle} \left(
{1 \over z_i^*}, {1 \over \bar{z}_i^*} \right) & \nameali{voptrans} \cr
& = ( {\rm phase \ factor} ) \times \left( {1 \over z_i^*} {1 \over
\bar{z}_i^*} \right)^2 \times {\cal V}_{\langle \lambda_i \vert}
\left( {1 \over z_i^*}, {1 \over \bar{z}_i^*} \right) \ , \cr }
$$
where the phase factor we pick up is minus one for space-time bosons and
$iY$ for space-time fermions. By eqs.~\outinv\ we pick up exactly the
same phase factor from vertex operators of the type ${\cal V}_{\langle
\lambda \vert}$. This amounts to an overall sign $(-1)^{N_B + N_{FP}}$,
$N_{FP}$ being the number of space-time fermion pairs and $N_B$
the number of space-time bosons. This sign exactly cancels the sign
produced by the $N_B + N_{FP}-2$ PCOs, which are anti-hermitean.
Finally, reordering the ghost factors in \Ttreeherm\ in accordance with
eq.~\Ttree , we obtain a minus sign cancelling the one that was
introduced by using eq.~\ghostherm.
Since the transformation factors $(z_i^*)^{-2} (\bar{z}_i^*)^{-2}$
appearing in eq.~\voptrans\ either cancels a similar one coming from the
ghost operators (for $i=1,2,3$), or is just the required jacobian to
transform ${\rm d}^2 z_i$ into ${\rm d}^2 \zeta_i$ where $\zeta_i = 1/z_i^*$ ($i
\geq 4$), we finally recover eq.~\Ttree\ multiplied by a phase factor
that, at the end, is just plus one.
This concludes the proof that our relation between
${\cal W}_{\langle \lambda \vert}^{(q)}$ and
${\cal W}_{\vert \lambda \rangle}^{(q)}$ leads to a hermitean
$T$-matrix at tree level away from the resonances.
\chapter{An explicit example}
In this section we provide an explicit example of the map \inoutq\
in the context of four-dimensional heterotic string models of the
Kawai-Lewellen-Tye (KLT) type [\Ref{KLT},\Ref{Anto}], where the internal
degrees of freedom are described by 22 left-moving and 9 right-moving
free complex fermions. We bosonize all these fermions (as well as the
four Majorana fermions $\psi^{\mu}$), using the explicit prescription
for bosonization in Minkowski space-time proposed in ref.~[\Ref{mink}].
In this formulation any state of the conformal field theory
(excluding the reparametrization ghosts) can be obtained by means of
non-zero mode creation operators from the generic ground state
which is specified by the
space-time momentum $k$, the ``momentum'' $J_0^{(L)} = {\Bbb A}_L$
of the 33 bosons $\Phi_{(L)}$ introduced by the bosonization,
and the superghost charge
$J_0^{(34)}=q={\Bbb A}_{34}$ which is (minus) the ``momentum''
of the field $\phi \equiv \Phi_{(34)}$ that is introduced when ``bosonizing''
the superghosts. Since $[ J_0^{(L)} , \Phi_{(K)} ] = \delta_K^{\ L}$,
the operator creating such a ground state
from the conformal vacuum is
$$ S_{\Bbb A} (z,\bar{z}) \ e^{i k \cdot X(z,\bar{z})} \ ,
\nfr{groundst}
where
$$ S_{\Bbb A} (z,\bar{z}) \equiv \prod_{L=1}^{34} e^{{\Bbb A}_L
\Phi_{(L)} (z,\bar{z})} \left( C_{(L)} \right)^{{\Bbb A}_L} \ ,
\nfr{spinfield}
is a spin field operator and $C_{(L)}$ is a cocycle factor, see
ref.~[\Ref{mink}] for details.
The range of values allowed for the ${\Bbb A}_L$ depends on the details
of the KLT model we happen to consider, see
refs.~[\Ref{KLT},\Ref{ammedm}]. We assume the level-matching condition
$L_0 - \bar{L}_0 = 0$ to be satisfied.
The hermitean conjugate of the operator $S_{\Bbb A} (z,\bar{z})$ can be
computed using the hermiticity properties of the various fields, as
outlined in Appendix B (see also ref.~[\Ref{mink}]). One finds
$$ \widehat{S}_{\Bbb A} (z,\bar{z}) = \left( \sigma_1^{(33)} {\Bbb
C}^{-1} \right)_{\Bbb A}^{\ \, \Bbb B} \ S_{\Bbb B} (z,\bar{z}) \ ,
\nfr{hermground}
where
$$ (\sigma_1^{(33)})_{{\Bbb A} {\Bbb B}} =
\left( \prod_{L=1}^{32} \delta_{{\Bbb A}_L,{\Bbb B}_L}\right) \ \delta_{{\Bbb
A}_{33} + {\Bbb B}_{33}, 0} \ \delta_{{\Bbb A}_{34},{\Bbb B}_{34}} \efr
and ${\Bbb C}^{-1}$ is the inverse of the ``charge conjugation matrix''
$$ {\Bbb C}_{{\Bbb A} {\Bbb B}} =
\left( \prod_{L=1}^{33} \delta_{{\Bbb A}_L + {\Bbb B}_L,0} \right) \
\delta_{{\Bbb A}_{34},{\Bbb B}_{34}} \ e^{ i \pi {\Bbb A} \cdot Y
\cdot {\Bbb B}}
\nfr{chargeconj}
defined in terms of the $34 \times 34$ cocycle matrix $Y_{KL}$ (see
refs.~[\Ref{mink},\Ref{ammedm}]).
The example we want to study is that of a physical
space-time fermion described by a ground state. To obtain a
BRST-invariant state one has to consider a vertex operator
which involves a linear combination of spin fields,
$$ {\cal V}^{(-1/2)}_{\vert {\Bbb V}, k \rangle} (z,\bar{z}) \ = \
{\kappa \over \pi} \ {\Bbb V}_{(-{1 \over 2})}^{\Bbb A} (k) \, S_{\Bbb A}
(z,\bar{z}) \ e^{i k \cdot X(z,\bar{z})} \ , \nfr{groundtwo}
where the spinor ${\Bbb V}^{\Bbb A}_{(-{1 \over 2})} (k)$ has superghost
charge $-1/2$, i.e. is proportional to
$\delta_{{\Bbb A}_{34},-1/2}$, and satisfies a Dirac equation which can
be obtained from the requirement that the $3/2$-order pole in the
operator product expansion (OPE) of the supercurrent $T_F^{[X,\psi]}$
with the operator \groundtwo\ vanishes.
If we define the gamma matrices by the OPE
$$ \psi^{\mu} (z) S_{\Bbb A} (w,\bar{w}) \ \buildchar{=}{\rm\scriptscriptstyle OPE}{ }\ {1 \over \sqrt{2}}
\left( {\bfmath \Gamma}^{\mu} \right)_{\Bbb A}^{\ \Bbb B} S_{\Bbb B}
(w,\bar{w}) {1 \over \sqrt{z-w}} + \ldots \ , \nfr{gammadef}
the Dirac equation assumes the matrix form
$$ ({\Bbb V}_{(-{1 \over 2})} (k))^T \ {\Bbb D} \, (k) = 0 \qquad {\rm or}
\qquad
\left( {\Bbb D}\, (k) \right)^T {\Bbb V}_{(-{1 \over 2})} (k) = 0 \ ,
\nfr{Diraceq}
where the Dirac operator is
$$ {\Bbb D}\, (k) = k_{\mu} {\bfmath \Gamma}^{\mu} - {\Bbb M} \ , \efr
${\Bbb M}$ being a mass operator that we do not need to write down
explicitly.
When the vertex operator is written as in eq.~\groundtwo\ we are no
longer free to choose the normalization of the spinor
${\Bbb V}_{(-{1 \over 2})} (k)$. It should be fixed in accordance
with eq.~\fermnormtwo .
In the next subsection we will explicitly verify that the
correct normalization is
$$ ({\Bbb V}_{(-{1 \over 2})} (k))^{\dagger} \
{\Bbb V}_{(-{1 \over 2})} (k) = \sqrt{2} \ |k^0| \ ,
\nfr{stringspinornorm}
which is analogous in structure to eq.~\ftspinornorm .
By using eq.~\hermground\ in the expression \inoutv\ we find the
``outgoing'' vertex operator corresponding to \groundtwo\ to be
$$ {\cal V}^{(-1/2)}_{\langle {\Bbb V}, k \vert} (z,\bar{z}) \ = \ -
\chi_{-1/2} {\kappa \over \pi} \left( {\Bbb V}^{{\Bbb A}}_{(-{1 \over 2})}
(k)\right)^* \left( \sigma_1^{(33)}
{\Bbb C}^{-1} \right)_{\Bbb A}^{\ \, {\Bbb B}} S_{\Bbb B}
(z,\bar{z}) \ e^{-i k \cdot X(z,\bar{z})} \ , \nfr{groundout}
where $\chi_{-1/2} = i Y$.
\section{A Sample Computation.}
We will now explicitly compute the amplitude for a space-time fermion
described by the vertex operator \groundtwo\ to absorb a zero-momentum
graviton. In particular we will obtain the relation \fivetwenty\ and show
how the sign $Y$ appearing in this formula is related to the choice of
cocycles.
Inserting eqs.~\groundtwo , \groundout\ and \gravvert\ into
eq.~\vmandvnone\ we obtain:
$$\eqalignno{T^0 ({\Bbb V},k \vert {\rm graviton} ; {\Bbb V}, k )
& = \ - C_0 \ \langle {\cal W}_{\langle {\Bbb V},k \vert}^{(-1/2)}
(z_1,\bar{z}_1) \ {\cal W}_{\vert {\rm grav} \rangle}^{(-1)}
(z,\bar{z}) {\cal W}_{\vert {\Bbb V},k \rangle}^{(-1/2)}
(z_2,\bar{z}_2) \rangle &\nameali{ampl} \cr
& =\ i \chi_{-1/2} C_0 \left( {\kappa \over \pi} \right)^3 \left( {\Bbb
V}^{\Bbb A}_{(-{1 \over 2})} (k) \right)^*
\left( \sigma_1^{(33)} {\Bbb C}^{-1} \right)_{\Bbb A}^{\ \, {\Bbb B}}
{\Bbb V}^{\Bbb C}_{(-{1 \over 2})} (k) \ \epsilon_{\mu}\ \times \cr
& \quad \langle S_{\Bbb B}(z_1,\bar{z}_1) \psi^{\mu}(z)
e^{- \Phi_{(34)} (z)} (C_{(34)})^{-1} S_{\Bbb C} (z_2,\bar{z}_2) \rangle
\ \times\cr
& \quad \langle \bar{\epsilon} \cdot \bar{\del} X(\bar{z}) e^{-i k \cdot
X(z_1,\bar{z}_1)} e^{i k \cdot X (z_2,\bar{z}_2)} \rangle \
\langle \bar{c} (\bar{z}_1) \bar{c} (\bar{z}) \bar{c} (\bar{z}_2)
c (z_1) c (z) c (z_2) \rangle \ . \cr }
$$
By explicit computation one finds
$$\eqalignno{ & \langle e^{-\Phi_{(34)} (z) } (C_{(34)})^{-1} \psi^{\mu}
(z) S_{\Bbb B}(z_1,\bar{z}_1) S_{\Bbb C} (z_2,\bar{z}_2) \rangle\ = &
\nameali{corr} \cr
&\qquad \ {1 \over \sqrt{2}} \left( {\bfmath \Gamma}^{\mu} \ {\Bbb C}_{(-1)}
\right)_{{\Bbb B}\, {\Bbb C}} { z_1 - z_2 \over (z-z_1)(z-z_2)} | z_1 -
z_2 |^{-2(2+m^2)} \ , \cr }
$$
where $m$ is the mass of the space-time fermion, $k^2 + m^2 = 0$, and we
introduced another family of ``charge conjugation matrices'' by
$$ \left( {\Bbb C}_{(q)} \right)_{{\Bbb A} {\Bbb B}} = \left(
\prod_{L=1}^{33} \delta_{{\Bbb A}_L + {\Bbb B}_L,0} \right) \
\delta_{{\Bbb A}_{34} +
{\Bbb B}_{34} + q + 2, 0 } \ e^{ i \pi {\Bbb A} \cdot Y \cdot {\Bbb B}}
\efr
for any value of $q \in {\Bbb Z}$.
Similarly one finds
$$\eqalignno{& \langle \bar{\epsilon} \cdot \bar\del X (\bar{z}) e^{-i k
\cdot X (z_1,\bar{z}_1)} e^{i k \cdot X (z_2,\bar{z}_2)} \rangle \ = \
i \bar{\epsilon} \cdot k {\bar{z}_1 - \bar{z}_2 \over (\bar{z} -
\bar{z}_1) (\bar{z} - \bar{z}_2) } | z_1 - z_2 |^{-2 k^2}\cr
& \langle \ \vert c(z_1) c(z) c(z_2) \vert^2 \ \rangle \ = \
\vert (z_1 - z) ( z- z_2) (z_1 - z_2) \vert ^2 \ . & \nameali{corrr} \cr }
$$
Substituting \corr\ and \corrr\ into eq.~\ampl\ we obtain
$$\eqalignno{ & T^0 ({\Bbb V}, k \vert {\rm graviton}; {\Bbb V}, k )\ = &
\nameali{ampltwo} \cr
&\ \chi_{-1/2} C_0 \left( {\kappa \over \pi} \right)^3 {1 \over
\sqrt{2} } \ \bar{\epsilon} \cdot k \
\epsilon_{\mu} \left( \left( {\Bbb V}_{(-{1 \over 2})} (k) \right)^{\dagger}
\sigma_1^{(33)} {\Bbb C}^{-1} {\bfmath \Gamma}^{\mu} {\Bbb C}_{(-1)}
{\Bbb V}_{(-{1 \over 2})}(k) \right) \ . \cr } $$
One may show that
$$ \left( {\bfmath \Gamma}^{\mu} {\Bbb
C}_{(-1)} \right)_{{\Bbb A} {\Bbb B}} \ = \
(-1)^{{\Bbb A}_{34}+1/2} \left( {\Bbb C}_{(-1)} {\bfmath \Sigma}
\left( {\bfmath \Gamma}^{\mu} \right)^T \right)_{{\Bbb A} {\Bbb B}} \ ,
\nfr{identityone}
where
$$ {\bfmath \Sigma}_{{\Bbb A} {\Bbb B}} \equiv \left( \prod_{L=1}^{34}
\delta_{{\Bbb A}_L,{\Bbb B}_L} \right) \
\exp \left\{ i \pi \sum_{L=1}^{33} Y_{34,L} {\Bbb B}_L \right\} \efr
and the sign $(-1)^{{\Bbb A}_{34}+1/2}$ is effectively equal to one,
since the matrices appearing in eq.~\ampltwo\ are sandwiched between
spinors with superghost charge $-1/2$. For the same reason the inverse
charge conjugation matrix ${\Bbb C}^{-1}$ is effectively equal to
$({\Bbb C}_{(-1)})^{-1}$.
Finally it is straightforward to verify that ${\bfmath \Gamma}^0$, as
defined by eq.~\gammadef , may also be written on the form
$$ {\bfmath \Gamma}^0 = i Y_{34,33} {\bfmath \Sigma} \sigma_1^{(33)}
\nfr{gammazero}
and since $\bfmath\Sigma$ and $\sigma_1^{(33)}$ anticommute,
$({\bfmath\Gamma}^0)^T = - {\bfmath\Gamma}^0$.
Inserting eqs.~\identityone\ and \gammazero\ into eq.~\ampltwo\ we
obtain
$$ \eqalignno{
& T^0 ({\Bbb V}, k \vert {\rm graviton}; {\Bbb V}, k )\ = & \numali \cr
& \qquad - i \chi_{-1/2} Y_{34,33} \ C_0 \left( {\kappa \over \pi} \right)^3
{1 \over \sqrt{2} } \ \bar{\epsilon} \cdot k \
\epsilon_{\mu} \left( \left( {\Bbb V}_{(-{1 \over 2})} (k) \right)^{\dagger}
({\bfmath \Gamma}^0)^T ({\bfmath \Gamma}^{\mu})^T
{\Bbb V}_{(-{1 \over 2})} (k) \right) \ . \cr } $$
At this point we may use the Gordon-like identity
$$
\left( {\Bbb V}_{(-{1 \over 2})}(k) \right)^{\dagger}
({\bfmath \Gamma}^0)^T ({\bfmath \Gamma}^{\mu})^T \
{\Bbb V}_{(-{1 \over 2})}(k) = - \sqrt{2} k^{\mu} \ . \efr
This equation can be proven directly using the Dirac equation \Diraceq ,
but it is easier to note that Lorentz covariance forces the right-hand
side to be proportional to $k^{\mu}$ and then fix
the proportionality constant by setting $\mu = 0$ and
using equation \stringspinornorm. Thus we finally obtain
$$ T^0 ({\Bbb V}, k \vert {\rm graviton}; {\Bbb V}, k ) \ = \
i \chi_{-1/2} Y_{34,33} \ C_0 \left( {\kappa \over \pi} \right)^3
\bar{\epsilon} \cdot k \
\epsilon \cdot k \ . \efr
This agrees with the correct result \rightres\ provided we choose
$$\chi_{-1/2} = i Y = i Y_{34,33} \efr
and shows that the sign $Y$ appearing in eq.~\fermnorm\ should be
identified with the component $Y_{34,33}$ of the cocycle matrix. At the
same time we have verified the correctness of the normalization
\stringspinornorm\ for the spinor ${\Bbb V}_{-{1 \over 2}} (k)$.
|
\section{1. Introduction and preliminaries.}
\vskip4pt plus2pt
{\bf 1.1.} {\it Introduction}
The algebraic approach to deformation-quantisation involves the
replacing of the
algebras of functions by non-commutative algebras.
In recent years we have seen a rapid developement of this approach
to quantisation, initiated by Drinfeld's [15]
realisation of Hopf algebras as deformations of Lie groups. Hopf
algebras are now commonly called quantum groups.
Quantum groups originated in the quantum inverse scattering method developed
by the Petersburg School and applied to quantisation of completely
integrable hamiltonian systems. Nowadays, however, it is believed
that quantisation-deformation and quantum groups in particular may
be applied to the description of spaces at the Planck scale.
Having this application in mind, it is important to develop
a kind of gauge theory involving quantum groups. Such a theory
was introduced by S. Majid and the author in [6]
in the framework of fibre bundles with quantum structure groups.
In this paper we review the main elements of the
quantum group gauge theory of [6].
The article is organised as follows.
In the remaining part of Section~1
we give a crash introduction to Hopf algebras
and non-commutative differential geometry. The reader familiar
with these topics may go directly to Section~2, where we describe
elements of the theory of quantum fibre bundles. Then in Section~3
we present gauge theory of such fibre bundles. We conclude
the paper with some remarks on other developments
of quantum group gauge theory and open problems in Section~4.
\vskip4pt plus2pt
{\bf 1.2.} {\it Hopf algebras.}
A unital algebra $H $ over a field $k$ is called a {\it Hopf algebra} if
there exist linear maps: a {\it coproduct} $\Delta :H\to H\otimes H$, a {\it
counit} $\epsilon :H\to k$
and an {\it antipode} ${\rm S}: H\to H$ which satisfy the following axioms [26]:
1. $(\Delta\otimes{\rm id})\circ\Delta = ({\rm id}\otimes\Delta)\circ\Delta$ ;
2. $({\rm id}\otimes\epsilon)\circ\Delta = (\epsilon\otimes{\rm id})\circ\Delta = {\rm id} $;
3. $m\circ({\rm id}\otimes{\rm S}) = m\circ({\rm S}\otimes{\rm id}) = 1\epsilon$.
Here and in what follows $m$ denotes the mulitpication map.
One should think of a Hopf algebra as a non-commutative generalisation of the
algebra of
regular functions on a group. In this case $\Delta$ corresponds to the group
multiplication and the axiom 1. states the associativity of this
multiplication.
Axiom 2. states the existence of the unit in a group and 3. is the existence
of inverses of group elements, written in a dual form. For this reason Hopf
algebras are also
called {\it quantum groups}.
For a
coproduct we use an explicit
expression $\Delta (a) = a{_{(1)}}\otimes a{_{(2)}}$, where the summation is
implied according to the Sweedler sigma convention [26], i.e.
$a{_{(1)}}\otimes a{_{(2)}} = \sum_{i\in I} a{_{(1)}}^ i\otimes a{_{(2)}}^ i$ for an
index set $I$. We also use the notation
$$
a{_{(1)}}{\otimes} a{_{(2)}}{\otimes}\cdots{\otimes} a_{(n)} =
(\Delta{\otimes}\underbrace{{\rm id}{\otimes}\cdots{\otimes}{\rm id}}_{n-2})\circ
\cdots\circ(\Delta{\otimes}{\rm id})\circ\Delta
$$
which describes a multiple action of $\Delta$ on $a\in H$.
A vector space $C$ with a coproduct $\Delta : C\to C\otimes C$
and the counit $\epsilon : C\to k$, satisfying axioms 1. and
2. is called a {\it coalgebra}.
A vector space $V$ is called a {\it right $H$-comodule} if there
exists a linear map $\rho_ R: V\to V{\otimes} H$, called a {\it right
coaction}, such that $(\rho_
R{\otimes}{\rm id})\circ\rho_ R = ({\rm id}{\otimes}\Delta)\circ\rho_ R$ and
$({\rm id}{\otimes} \epsilon )\circ\rho_ R = {\rm id}$.
We say that a unital algebra $P$ over $k$ is a {\it right H-comodule
algebra} if $P$ is a right $H$-comodule with a coaction
$\Delta_ R :P\to P\otimes H$, and $\Delta_ R$ is an algebra
map. The algebra structure of $P{\otimes} H$ is that of a tensor product
algebra. For a coaction $\Delta_ R$ we use an explicit notation
$\Delta_ R u = u_{(0)}\otimes u_{(1)}$, where
the summation is also implied. Notice that $u_{(0)}\in
P$ and $u_{(1)}\in H$. If $P$ is a right $H$-comodule so is
$P\otimes P$ with a coaction $\Delta_ R$
$$
\Delta_ R(u\otimes v) = u{_{(0)}} \otimes
v{_{(0)}}\otimes u{_{(1)}} v{_{(1)}}. \eqno{(1)}
$$
If $P$ is a right $H$-comodule algebra then $P^{coH}$ denotes a fixed
point subalgebra of $P$, i.e.
$P^{coH} = \{u\in P :\Delta_ R u = u\otimes 1\}$.
$P^{coH}$ is a subalgebra of $P$ with a natural inclusion $j:
P^{coH} \hookrightarrow P$ which we do not write explicitly later on.
Let $H$ be a Hopf algebra, $B$ be a unital algebra over $k$, and
let $f,g :H\to B$ be linear maps. A {\it convolution product} of
$f$ and $g$ is a linear map $f*g: H\to B$ given by
$(f*g)(a) = f(a{_{(1)}})g(a{_{(2)}})$, for any $a\in H$. With respect to
the convolution product, the set of all linear
maps $H\to B$ forms an
associative algebra with the unit $1\epsilon$. We say that a
linear map $f: H\to B$ is
{\it convolution invertible} if there is a map $f^{-1}: H\to
B$ such that
$f*f^{-1} = f^{-1}*f = 1\epsilon$. The set of all convolution
invertible maps $H\to B$ forms a multiplicative group. Similarly if
$V$ is a right $H$-comodule and $f:V\to B$, $g:H\to B$ are linear maps
then we define a convolution product $f*g:V\to B$ to be $(f*g)(v) =
f(v{_{(0)}})g(v{_{(1)}})$.\vskip4pt plus2pt
{\bf 1.3.} {\it Differential structures.}
Let $P$ be a unital algebra over $k$. Denote by ${\Omega^{1}P}$ the $P$-bimodule
${\rm ker} m$,
where $m: P\otimes P \to P$ is a multiplication map. Let ${\rm d_ U} : P\to {\Omega^{1}P}$ be
a
linear map
$$
{\rm d_ U} u = 1\otimes u - u\otimes 1. \eqno{(2)}
$$
It can be easily checked that ${\rm d_ U}$ is a differential, known as the Karoubi
differential. We call the pair $({\Omega^{1}P}, {\rm d_ U})$ the {\it universal differential
structure on P} [19, 20]. ${\Omega^{1}P}$ should be understood as a bimodule
of 1-forms. We say that $ (\Omega^ 1(P) ,{\rm d})$
is a {\it first order differential calculus} on $P$ if there exists
a subbimodule ${\cal N}\subset{\Omega^{1}P}$ such that $ {\Omega^{1}(P) } = {\Omega^{1}P} /{\cal N}$ and
${\rm d} = \pi\circ{\rm d_ U}$, where $\pi: {\Omega^{1}P}\to{\Omega^{1}(P) }$ is a canonical projection.
It is then said that $({\Omega^{1}(P) }, {\rm d})$ is generated by ${\cal N}$. Let a differential
structure $(\Omega^ 1(H), {\rm d})$ on a Hopf algebra $H$ be generated by
${\cal N} \subset\Omega^ 1H$. We say that $(\Omega^ 1(H), {\rm d})$
is a {\it bicovariant differential calculus} [29] if there exists a unique
right ideal ${\cal Q}\subset{\rm ker}\epsilon$ such that $H\otimes{\cal Q} =\kappa({\cal N})$,
where $\kappa :H\otimes H\to H\otimes H$, $\kappa : a\otimes b \mapsto
ab{_{(1)}}\otimes b{_{(2)}}$, and ${\rm Ad_{R}}({\cal Q}) \subset {\cal Q}\otimes H$, where ${\rm Ad_{R}} : H\to
H\otimes H$
is a right adjoint coaction
$$
{\rm Ad_{R}} :a\mapsto a{_{(2)}}\otimes ({\rm S} a{_{(1)}})a_{(3)}.\eqno{(3)}
$$
The universal
differential envelope is the unique differential algebra $(\Omega P, {\rm d})$
containing $({\Omega^{1}P}, {\rm d_ U})$ as its 1-st order part.
\section{2. Fibre bundles.}
In this section we report the basic elements of the theory of
quantum fibre bundles of S. Majid and the author
[6]. The detailed analysis of quantum
group gauge theory on classical spaces may be found in
[7]. All the algebras are over a field $k$ of complex
or real numbers. Except for Section~2.4 and
Example~3.1.4
we work with the universal differential structure.
\vskip4pt plus2pt
{\bf 2.1.} {\it Quantum principal bundles.}
Let $H$ be a Hopf algebra, $P$ a right $H$-comodule algebra
with a coaction
$\Delta_ R :P\to P\otimes H$. We define a canonical map $\chi :P\otimes P \to
P\otimes H$,
$$
\chi = (m\otimes {\rm id})\circ ({\rm id} \otimes \Delta_ R).
\eqno{(4)}
$$
Explicitly, $\chi (u\otimes v) = uv{_{(0)}}\otimes
v{_{(1)}}, $ for any $u,v\in P$. We say that the coaction
$\Delta_ R$ is
{\it free} if $\chi$ is a surjection and it is {\it exact}
if ${\rm ker} \chi = P({\rm d} P^{coH})P,$ where ${\rm d}$ denotes the
universal differential (2) and $P^{coH}$ is a fixed point subalgebra
of $P$. We denote $P({\rm d} P^{coH})P$ by ${\Omega^{1}P}_{\rm hor}$ and call
its elements {\it horizontal forms}. Although the freeness and exactness
conditions are algebraic in this formulation one should notice
that in fact the latter one is a condition on differential
structures on $P$ and $P^{coH}$. This becomes clear
in Section~2.4.
The map $\chi\mid_{\Omega^{1}P}$ has a natural geometric
interpretation as a dual to the map $ {\cal G}\to T_ uX$, which to
each element of the Lie algebra $\cal G$ of a group $G$
associates a fundamental vector field on a manifold $X$ on which $G$
acts.
\defin{Definition}{2.1.1 Let $H$ be a Hopf algebra, $(P ,\Delta_{R})$ be a
right $H$-comodule
algebra and let $B = P^{coH}$. We say that $P(B,H)$ is a {\it quantum
principal bundle} within the differential
envelope, with a structure quantum group $H$ and a base $B$ if the
coaction $\Delta_ R$ is free and exact.}
This definition reproduces the classical situation (but in a dual
language) in which a group $G$ acts freely on a total space $X$ from
right, and a base manifold $M$ is defined as $M = X/G$. The freeness
of the action of $G$ on $X$ means that a map
$X\times G \to X\times X$, $(u,g)\mapsto (u,ug)
$ is an inclusion. In the classical situation and the commutative
differential structure the exactness follows from the
freeness. This
is no longer true in a non-commutative extension.
The notion of a quantum principal bundle is strictly related to
the theory of algebraic extensions [25] since $P(B,H)$
is a Hopf-Galois extension
of $B$ to $P$ by a Hopf algebra $H$.
Yet another way of defining of a quantum principal bundle makes use
of the notion of a {\it translation map}, which proves very useful
in analysis of the structure of quantum bundles [4].
{_{(3)}}{Proposition}{2.1.2.}{
Let $H$ be a Hopf algebra, $P$ a right
$H$-comodule algebra and $B = P^{co H}$.
Assume that the coaction $\Delta_ R$ is free.
Then $P(B,H)$ is a quantum principal bundle iff there exists a linear map
$\tau:H\to
P\otimes{}_ B P$, given by $\tau(a) = \sum_{i\in I}u_
i\otimes{}_ B v_ i$, where $\sum_{i\in I}u_ i\otimes v_ i \in
\chi^{-1}(1\otimes a)$. The map $\tau$ is called a {\it translation map}.}
A translation map is a well-known object in the classical bundle theory
[18, Definition~2.1].
Classically, if $X$ is a manifold on which a Lie group $G$ acts
freely then the translation map $\hat{\tau}:X\times{}_ MX\to G$, where
$M=X/G$, is defined by $u\hat{\tau}(u,v)=v$.
Dualising this construction we arrive immediately at
the map $\tau$ above.
\vskip4pt plus2pt
{\bf 2.2.} {\it Examples of quantum principal bundles}
\remar{Example\ {2.2.1.}\ }
{{\it A trivial quantum principal bundle.}
Let $H$ be a Hopf algebra, $P$ a right $H$-comodule
algebra and $B = P^{coH}$.
Assume there is a convolution invertible map $\Phi :
H\rightarrow P$ such that
$
\Delta_ R\Phi = (\Phi\otimes{\rm id})\Delta ,\quad \Phi(1)=1 ,
$
i.e.~$\Phi$ is an intertwiner. Then $P(B,H)$ is a quantum principal
bundle called a {\it trivial quantum principal bundle} and denoted by
$P(B,H,\Phi)$. The word {\it trivial} refers to the fact that $P
\cong B\otimes H$ as vector spaces with an isomorphism $\Theta_\Phi
: P\to B\otimes H$, $\Theta_\Phi:
u\mapsto u{_{(0)}}\Phi^{-1}(u{_{(1)}})\otimes u{_{(2)}}$. Moreover, as algebras
$P\cong B_\Phi\# H$, where
${}_\Phi\#$ denotes a crossed product [1], with the
isomorphism $\Theta_\Phi$
above. Explicitly, the product in
$B_\Phi\# H$ is given by
$$
(b_ 1\otimes a^ 1)(b_ 2\otimes a ^ 2) = b^ 1\Phi( a^1{_{(1)}}) b_2
\Phi(a^2{_{(1)}})\Phi^{-1}(a^1{_{(2)}} a^2{_{(2)}})\otimes a^1_{(3)} a^2_{(3)} .
$$
Such an algebra $P$ is also known as a {\it cleft extension} of $B$
[27, 14].
The map $\tau = (\Phi^{-1}\otimes_ B\Phi)\circ\Delta$ is a translation
map in $P(B,H,\Phi)$.}
For a trivial quantum principal bundle $P(B,H,\Phi)$ we define a
{\it gauge transformation} as a convolution invertible map $\gamma:
H\to B$ such that $\gamma(1) =1$. The set of all gauge transformations
of $P(B,H,\Phi)$ forms a group with respect to the
convolution product. This group is denoted by ${\cal H}(B)$. Gauge
transformations relate different trivialisations of $P(B,H,\Phi)$:
$\Psi :H\to P$ is a trivialisation of $P(B,H,\Phi)$ iff there exists $\gamma
\in {\cal H}(B)$ such that $\Psi = \gamma*\Phi$. They also
have a clear meaning in the theory of crossed products. The
following proposition is a special case of the result of Doi [13]
(see also [21, Proposition 4.2]).
{_{(3)}}{Proposition}{2.2.2.}{
Let $P(B,H,\Phi)$ be a trivial quantum principal bundle. Let for any
trvialisation $\Psi$ of $P(B,H,\Phi)$, $\Theta_\Psi :
B_\Psi\# H\to B_\Phi\# H$ be a crossed product algebra
isomorphism such that $\Theta_\Psi\mid_ B = {\rm id}$ and $\Delta_ R\Theta
_\Psi = (\Theta_\Psi\otimes{\rm id})\Delta_ R$. Then there is a
bijective correspondence between all isomorphisms $\Theta_\Psi$
corresponding to all trivialisations $\Psi$ and the gauge transformations
of $P(B,H,\Phi)$.}
\remar{Example\ {2.2.3.}\ }{
{\it Quantum principal bundle on a quantum homogeneous space.}
Let $H$ and $P$ be Hopf algebras. Assume, there is a Hopf algebra
projection $\pi : P\rightarrow H$. Define a right coaction of $H$ on
$P$ by
$
\Delta_ R = ({\rm id}\otimes\pi)\Delta : P\rightarrow P\otimes H.
$
Then $B = P^{coH}$ is a {\it quantum quotient space}, a special
case of a {\it quantum homogeneous space}.
Assume that ${\rm ker}\pi \subset m\circ ({\rm ker}\pi\mid_ B\otimes P)$. Then
$P(B,H)$ is a quantum principal bundle within the differential
envelope. This bundle is denoted by $P(B,H,\pi)$.
The translation map $\tau: H\to
P\otimes{}_ B P$ in $P(B,H,\pi)$ is given by
$\tau(a) = {\rm S} u{_{(1)}}\otimes{}_ B u{_{(2)}}$,
where $u\in\pi^{-1}(a)$.}
A large number of examples of quantum bundles on quantum homogeneous
spaces has been found in [22]. The simplest
and probably the most fundamental one is
\remar{Example\ {2.2.4.}\ }
{{\it The quantum Hopf fibration \rm [6, Section~5.2]}. The
total space of this bundle is the quantum group $SU_
q(2)$, as an algebra generated by the identity and a matrix
$T = (t_{ij}) = \pmatrix{\alpha & \beta \cr \gamma &\delta}$, subject
to the homogeneous relations
$$
\alpha\beta = q\beta \alpha , \quad \alpha\gamma = q
\gamma\alpha , \quad \alpha \delta = \delta\alpha +
(q-q^{-1})\beta\gamma ,\quad
\beta\gamma = \gamma\beta, \quad \beta\delta = q\delta \beta
,\quad \gamma \delta = q \delta \gamma ,
$$
and a determinant relation
$\alpha\delta-q\beta\gamma=1$, $q\in k^*$. $SU_ q(2)$ has a matrix quantum
group
structure,
$$
\Delta t_{ij} = \sum_{k =1}^ 2 t_{ik}{\otimes} t_{kj}, \quad
{\epsilon}(t_{ij})=\delta_{ij},\quad {\rm S} T = \pmatrix{\delta &-q^{-1}
\beta \cr -q\gamma & \alpha}.
$$
The structure quantum group of the quantum Hopf bundle is an algebra
of functions on $U(1)$, i.e. the algebra $k[Z,Z^{-1}]$ of formal
power series in $Z$ and $Z^{-1}$, where $Z^{-1}$ is an inverse of
$Z$. It has a standard Hopf algebra structure
$$
\Delta Z^{\pm 1} = Z^{\pm 1}{\otimes} Z^{\pm 1} , \quad
{\epsilon}(Z^{\pm 1}) = 1, \quad
{\rm S} Z^{\pm 1} = Z^{\mp 1} .
$$
There is a Hopf algebra projection $\pi: SU_ q(2) \to
k[Z,Z^{-1}]$,
$$
\pi : \pmatrix{\alpha & \beta \cr \gamma &\delta} \mapsto \pmatrix{Z &
0\cr 0 & Z^{-1}},
$$
which defines a right coaction $\Delta_ R : SU_ q(2) \to SU_
q(2) {\otimes} k[Z,Z^{-1}]$ by $\Delta_ R =({\rm id}{\otimes}\pi)\circ\Delta$.
Finally $S_ q^ 2\subset SU_ q(2)$ is a quantum two-sphere
[24], defined as a fixed point subalgebra, $S_ q^ 2 = SU_
q(2)^{cok[Z,Z^{-1}]}$. $S_ q^ 2$ is generated by $\{1, b_ - =
\alpha\beta , b_ + =\gamma\delta ,b_ 3 = \alpha\delta\}$ and the
algebraic relations in $S_ q^ 2$ may be deduced from those in
$SU_ q(2)$.
It was shown in [6] that $SU_
q(2)(S_ q^ 2,k[Z,Z^{-1}],\pi)$ is a non-trivial quantum
principal bundle over the homogeneous space.}
The other examples of quantum principal bundles constructed in [22]
include:
$$
U_ q(n)(S_ q^{2n-1}, U_ q(n-1),\pi),$$
$$
SU_ q(n)(S_ q
^{2n-1}, SU_ q(n-1),\pi),$$
$$SU_ q(n)({\bf C}{\bf P}_ q^{n-1}, U_ q(n-1),
\pi),$$
$$
U_ q(n)(G_ k({\bf C}^ n_ q), U_ q(k)\otimes U_ q(n-k), \pi),
$$
where $G_ k({\bf C}^ n_ q)$ is a quantum Grassmannian.
\remar{Remark\ {2.2.5.}\ }{The quantum sphere $S^2_q$ considered
in Example~2.2.4. is the special case of the most general
quantum sphere $S^2_q(\mu ,\nu)$, where $\mu\neq\nu$ are
real parameters such that $\mu \nu \geq 0$ (see [24] for
details). Precisely $S^2_q = S^2_q(1,0)$. It can
be shown that $S^2_q$ is the only quantum sphere
which can be interpreted as a quotient space of $SU_q(2)$
by $k[Z,Z^{-1}]$ in the sense
of Example~2.2.3. It turns out, however, that $S^2_q(\mu ,\nu)$
may be veiwed as a quotient space of $SU_q(2)$ by
a {\it coalgebra} $C = SU_q(2)/J$, where $J$ is a right
ideal in $SU_q(2)$ generated by
$$
p(q\alpha^2 - \beta^2) + \alpha\beta -pq, \quad
p(q\gamma^2 - \delta^2) + \gamma\delta + p, \quad
p(q\alpha\gamma - \beta\delta) + q\beta\gamma ,
$$
where $p = \sqrt{\mu\nu}/(\mu-\nu)$ [5]. Precisely
$$
S^2_q(\mu,\nu) = \{u\in SU_q(2);\;\; u{_{(1)}}\otimes\pi(u{_{(2)}}) =
u\otimes\pi( 1)\},
$$
where $\pi:SU_q(2)\to C$ is the canonical surjection.
It can be shown that the vector space $C$ is spanned
by $1=\pi(1)$, $x_ n = \pi(\alpha^n)$ and
$y_n = \pi(\delta^n)$ (cf. definition of $\pi$ in Example~2.2.4).
One would like to view $SU_q(2)$ as a total space of a quantum
principal bundle over $S_q^2(\mu, \nu)$similarly as in Example~2.2.4. Since
$C$ is not a Hopf algebra one needs to generalise the notion
of a bundle. In [5] we proposed the following generalisation
of Definition~2.1.1. Let $C$ be a coalgebra and let $P$ be an algebra
and a right $C$-comodule. Assume that there
is an action $\rho : P\otimes C\otimes P \to P\otimes
C$ of $P$ on $P\otimes C$ and an element $ 1\in C$
such that $\Delta_R\circ m = \rho\circ(\Delta_R\otimes{\rm id})$ and
for any $u, v\in P$, $\rho(u\otimes 1 , v)
= \chi(u\otimes v)$. Then $B = \{ u\in P;\;\; \Delta_R u =u\otimes 1\}$
is a subalgebra of $P$, and we say that $P(B,C,\rho)$
is a {\it quantum $\rho$-principal bundle} over $B$
if the coaction $\Delta_R$ is free and exact.
In the above example of the quantum sphere $S^2_q(\mu,\nu)$ the action
$\rho$ is given by $\rho(u\otimes c, v) = uv{_{(1)}}\otimes
\rho_0(c,v{_{(2)}})$, where $\rho_0$ is a natural right
action of $SU_q(2)$ on $C$.}
\vskip4pt plus2pt
{\bf 2.3.} {\it Quantum associated bundles}
\defin{Definition} {2.3.1.
\rm Let $P(B,H)$ be a quantum principal bundle and let $V$ be a right
$H^{\rm op}$-comodule algebra, where $H^{\rm op}$ denotes the algebra
which is isomorphic to $H$ as a vector space but has an opposite
product, with coaction $\rho_{R} : V \rightarrow V
\otimes H$. The space $P \otimes V$ is naturally endowed with a right
$H$-comodule structure $\Delta_{E} : P \otimes V
\rightarrow P \otimes V \otimes H$ given by
$\Delta_{E} (u \otimes v) = u{_{(0)}} \otimes
v{_{(0)}} \otimes u{_{(1)}} v{_{(1)}}$
for any $u \in P$ and $v \in V$. We say that the fixed point subalgebra
$E$ of $P{\otimes} H$ with respect to $\Delta_ E$
is a {\it quantum fibre bundle associated to $P(B,H)$} over $B$ with
structure quantum group $H$ and standard fibre $V$. We denote it by $E =
E(B,V,H)$.}
It can be easily shown that $B$ is a subalgebra of $E$ with the inclusion
$j_ E = b{\otimes} 1$. The inclusion $j_ E$ provides $E$ with the
structure of a left $B$-module.
\remar{Example\ {2.3.2.}\ }
{ Let $P(B,H,\Phi)$ be a trivial quantum principal bundle and let $V$ be
as in Definition~2.3.1. Assume also that $H$
has a bijective antipode. The associated bundle $E(B,V,H)$ is called
a {\it trivial quantum fibre bundle}. Trivialisation $\Phi :H\to P$ induces a
map $\Phi_ E :V\to E$,
$
\Phi_{E} (v) = \sum \Phi ({\rm S}^{-1} v{_{(1)}}) \otimes
v{_{(0)}}$
which allows one to identify $E$ with $B{\otimes} V$ as vector spaces via
the linear isomorphism $b{\otimes} v\mapsto b\Phi_ E(v)$. As an algebra,
$E$ is isomorphic to a certain crossed product algebra $B\# V$
[2]. }
The following proposition shows that a
quantum principal bundle is a fibre bundle associated to itself.
{_{(3)}}{Proposition}{2.3.3.} {
A quantum principal bundle $P(B,H)$ is a fibre bundle associated to
$P(B,H)$ with the fibre which is isomorphic to $H$ as an algebra and
with the coaction $\rho_ R=({\rm id}\otimes {\rm S})\circ\Delta'$, where
$\Delta'$ denotes the opposite coproduct, $\Delta'(a) = a{_{(2)}}{\otimes} a{_{(1)}}$,
for any $a\in H$.}
{}From the point of view of a gauge theory it is important to consider
cross-sections
of a vector bundle. In this algebraic setting a cross-section is defined as
follows
\defin{Definition}{ 2.3.4.
Let $E(B,V,H)$ be a quantum fibre bundle associated to a quantum
principal bundle $P(B,H)$. A left $B$-module map
$s: E\to B$ such that $s(1) = 1$ is called a {\it cross section} of
$E(B,V,H)$. The set of cross sections of $E(B,V,H)$ is denoted by
$\Gamma(E)$. }
{_{(3)}}{Lemma}{ 2.3.5.} {
If $s:E\to B$ is a cross section of a quantum fibre bundle $E(B,V,A)$ then
$s\circ j_ E = {\rm id}$.}
The result of trivial Lemma~2.3.5. justifies the term
cross section used in Definition~2.3.4. We remark that the
definition of a cross section of a quantum fibre bundle analogous to the
one we use
here was first proposed in [17].
We analyse cross-sections more closely in Section~3.3.
\vskip4pt plus2pt
{\bf 2.4.} {\it Quantum principal bundles with general differential
structures.}
The detailed analysis of quantum principal bundles with general
differential structures goes far beyond the scope of this
paper. Here we give only a definition of a quantum principal bundle
with general differential structure. We refer the interested reader to the
fundamental
paper [6]. More explicit exposition may be
also found in [2].
Let $({\Omega^{1}(P) } ,{\rm d})$ be a first order differential calculus on a right $H$-comodule
algebra $P$ generated
by ${\cal N}\subset{\Omega^{1}P}$ and let $(\Omega^ 1(H), {\rm d})$ be a bicovariant
differential structure on $H$ generated by the right ideal ${\cal Q}\subset
{\rm ker}\epsilon$. We say that differential structures $({\Omega^{1}(P) }, {\rm d})$
and $(\Omega^ 1(H),{\rm d})$ {\it agree} with each other if
$\Delta_ R({\cal N})\subset{\cal N}\otimes H$ , where $\Delta_ R$ is given
by (1), and $\chi({\cal N})\subset P\otimes{\cal Q}$.
If differential structures on $P$ and $H$ agree we can define
a map $\chi_{\cal N} :{\Omega^{1}(P) }\to P\otimes{\rm ker}\epsilon /{\cal Q}$ as follows.
Let $\pi_{\cal N} :{\Omega^{1}P}\to{\Omega^{1}(P) }$ and $\pi_{\cal Q} :{\rm ker}\epsilon\to
{\rm ker}\epsilon /{\cal Q}$ be canonical projections. Then for any $\rho\in{\Omega^{1}(P) }$
take any $\rho_ U\in\pi_{\cal N}^{-1}(\rho)$ and define
$\chi_{\cal N}(\rho) = ({\rm id}\otimes\pi_{\cal Q})\circ\chi(\rho_ U)$, where
$\chi$ is a canonical map (4). We say that the coaction
$\Delta_ R:P\to P\otimes H$ is {\it exact} with respect to differential
structures generated by ${\cal N}$ and ${\cal Q}$ if ${\rm ker}\chi_{\cal N} =
P\Omega^ 1(P^{coH}) P$. Finally we define a quantum principal
bundle with $P(B,H)$ with differential structure generated by ${\cal N}$
and ${\cal Q}$ if the coaction $\Delta_ R$ is free and exact with respect
to this structure.
\section{3. Gauge Theory.}
In this section we analyse more closely the
structure of quantum bundles. We introduce
the formalism of connections and take a closer look at cross sections and
gauge transformations in general (non-trivial) quantum bundles.
\vskip4pt plus2pt
{\bf 3.1. }{\it Connections = gauge fields.}
{}From the point of view of gauge theories principal connections
are the gauge fields. In the definition of a principal connection an important
r\^ole is played by a right adjoint coaction of $H$ on itself (3).
Since ${\rm Ad_{R}} ({\rm ker} \epsilon)\subset{\rm ker}\epsilon\otimes H$,
we can define a coaction $\Delta_ R
:P\otimes{\rm ker}\epsilon \to P\otimes{\rm ker}\epsilon\otimes H$ by
$
\Delta_ R (u\otimes a) = u{_{(0)}}\otimes a{_{(2)}}\otimes
u{_{(1)}}({\rm S} a{_{(1)}})a_{(3)}.
$The canonical map $\chi :{\Omega^{1}P}\to P\otimes{\rm ker}\epsilon$
is equivariant, i.e.
$
\Delta_ R \chi = (\chi\otimes {\rm id} )\Delta_
R,
$
where $\Delta_ R$ on ${\Omega^{1}P}$ is given by (1).
{}From the definition of a quantum principal bundle we deduce
that the following sequence
$$
0\to\Gamma_{hor}\buildrel{j}\over{\to}{\Omega^{1}P}\buildrel{\chi}\over\to
P\otimes{\rm ker}\epsilon\to 0
$$
is an exact sequence of equivariant maps. A connection in
$P(B,H)$ is a right-invariant splitting of this sequence. In
other words, if there is a map
$
\sigma
:P\otimes{\rm ker}\epsilon\to{\Omega^{1}P}
$
such that $\Delta_
R\sigma = (\sigma\otimes{\rm id})\Delta_{R}$ and
$\chi\circ\sigma = {\rm id}$, then a connection in
$P(B,H)$ is identified with a linear projection $\Pi: {\Omega^{1}P}\to{\Omega^{1}P}$, $\Pi =
\sigma\circ\chi\mid_{\Omega^{1}P}$. Obviously, $\Delta_ R\Pi =
(\Pi\otimes{\rm id})\Delta_ R$. The connection $\Pi$ is {\it strong} if and only if
$({\rm id} -\Pi){\rm d} P \subset \Omega^ 1BP$, [17].
We denote ${\Omega^{1}P}_{\rm ver} = {\rm Im}\Pi$. Every
$\alpha\in{\Omega^{1}P}_{\rm ver}$ is said to be a {\it vertical
1-form}. If there is a connection in $P(B,H)$, then
${\Omega^{1}P} = {\Omega^{1}P}_{\rm hor}\oplus{\Omega^{1}P}_{\rm ver}$.
Next we define a map $\omega :H\to {\Omega^{1}P}$, by
$$
\omega(a) = \sigma(1\otimes (a-\epsilon(a))).
$$
The map $\omega$ is called a {\it connection 1-form} of the
connection $\Pi$.
{_{(3)}}{Theorem} {3.1.1.} {Let $P(B,H)$ be a quantum principal bundle and let
$\Pi$
be a connection in
$P(B,H)$. A connection form $\omega$ has the following properties:
1. $\omega(1) = 0$;
2. $\forall\; a\in H, \quad\chi\omega (a) = 1\otimes
(a- \epsilon (a))$;
3. $\Delta_{R} \circ\omega = (\omega\otimes{\rm id})\circ{\rm Ad_{R}}$.
Conversely, if $\omega :H\to {\Omega^{1}P}$ is a linear map
obeying 1-3, then $\Pi = m\circ ({\rm id}\otimes\omega)\chi\mid_{\Omega^{1}P}$ is
a connection with a connection 1-form $\omega$.}
Having a connection $\Pi$ in a quantum principal bundle $P(B,H)$ one can
define
the horizontal projection as a complimentary part of $\Pi$, and a covariant
derivative
as a horizontal part of ${\rm d}$ (for details see [6]). As a result one
defines a curvature of a strong connection $\omega$ as $F =
{\rm d}\omega+\omega*\omega$
[17].
\remar{Example\ {3.1.2.}\ }{
{\it Strong connection in a trvial bundle.}
Let $P(B,H,\Phi)$ be a trivial quantum principal bundle as
before, and let $\beta :H\to \Omega^ 1B$ be any linear
map such that $\beta(1)=0$. Then the map
$
\omega = \Phi^{-1}*\beta*\Phi +\Phi^{-1}*{\rm d}\Phi
$
is a connection 1-form in $P(B,H,\Phi)$. Its curvature is easily computed
to be $F = \Phi^{-1}*({\rm d}\beta +\beta*\beta)*\Phi$.}
\remar{Example\ {3.1.3.}\ }
{{\it Canonical connection. } Let $P(B,H,\pi)$ be a quantum principal bundle
over the
homogeneous space $B$ as described in
Example~2.2.3. Assume, there is an algebra
inclusion $i :H\hookrightarrow P$ such that $\pi\circ i ={\rm id}$,
$\epsilon_ P(i(a)) = \epsilon_ H (a)$, for any $a\in H$ and such
that
$
({\rm id}\otimes\pi) {\rm Ad_{R}} i = (i\otimes {\rm id}) {\rm Ad_{R}} .
$
Then the map $\omega(a) = {\rm S} i(a){_{(1)}}{\rm d} i(a){_{(2)}}$
is a connection 1-form in $P(B,H,\pi)$. This connection
is strong if $i$ is an
intertwiner for the right coaction [2, Lemma~5.5.5].}
\remar{Example\ {3.1.4.}\ }
{{\it The Dirac $q$-monopole.}
Consider the quantum Hopf fibration of Example~2.2.4.
Let a differential structure $(\Omega^ 1(SU_ q(2)), {\rm d})$ be given by the
3D caluclus of Woronowicz [28]. $\Omega^ 1(SU_ q(2))$
is generated by the
forms $\omega^ 0 = \delta{\rm d}\beta - q^{-1}\beta{\rm d}\delta$,
$\omega^ 1 = \delta{\rm d}\alpha - q^{-1}\beta{\rm d}\gamma$,
$\omega^ 2 = \gamma{\rm d}\alpha - q^{-1}\alpha{\rm d}\gamma$
and the
relations
$$
\omega^0\alpha = q^{-1}\alpha\omega^0 , \quad \omega^0\beta =
q\beta\omega^0 ,\quad
\omega^1\alpha = q^{-2}\alpha\omega^1 ,
$$
$$
\omega^1\beta =
q^2\beta\omega^1, \quad
\omega^2\alpha = q^{-1}\alpha\omega^2 , \quad \omega^2\beta =
q\beta\omega^2 .
$$
The remaining relations can be obtained by the replacement
$\alpha\rightarrow\gamma$, $\beta\rightarrow\delta$. One can show
that $SU_ q(2)(S_ q^ 2, k[Z,Z^{-1}], \pi)$ is a quantum
principal bundle with this differential structure. We define
the connection one form $\omega:k[Z,Z^{-1}]\to\Omega^{1}(SU_ q(2))$
by
$$
\omega(Z^ n) = {{q^{-2n} - 1}\over{q^{-2} -1}}\omega^ 1.
$$
In [6] it has been shown that $\omega$ is a canonical connection in
$SU_ q(2)(S_ q^ 2, k[Z,Z^{-1}], \pi) $ which reduces to
the Dirac monopole of charge 1 [16] when $q\to 1$. The curvature
of $\omega$ is $F(Z^ n) = {{q^{-2n} - 1}\over{q^{-2} -1}}\omega^ 0
\wedge\omega^ 2$. The q-deformed Dirac monopole of any charge is
discussed
in [12].}
\vskip4pt plus2pt
{\bf 3.2.} {\it Cross sections = matter fields}
In this section we use the notion of a translation map in a quantum
principal bundle $P(B,A)$ to identify cross sections of a quantum
fibre bundle $E(B,V,A)$ with equivariant maps $V\to P$. In gauge theories
such maps play a role of matter fields. Recall that a
linear map $\phi:V\to P$ is said to be equivariant if $\Delta_ R
\phi = (\phi{\otimes}{\rm id})\rho_ R$, where $\rho_ R$ is a right coaction
of $A$ on $V$. In particular, our identification
implies that a quantum principal bundle is trivial if it admits a
cross section which is an algebra map.
{_{(3)}}{Theorem} {3.2.1} {
Let $H$ be a Hopf algebra with a bijective antipode. Cross sections of
a quantum fibre bundle $E(B,V,H)$ associated to a
quantum principal bundle $P(B,H)$ are in bijective
correspondence
with equivariant maps $\phi :V\to P$ such that $\phi(1) = 1$.}
\Proof A map $\phi :V \to P$ induces a
cross section $s$ of $E(B,V,H)$, by $s =
m\circ({\rm id}{\otimes}\phi)$. Conversely, for any $s\in\Gamma(E)$ we define a
map $\phi: V\to P$ by
$$
\phi : v\mapsto {\tau^{(1)}}}\def\taut{{\tau^{(2)}}({\rm S}^{-1}v{_{(1)}})s(\taut({\rm S}^{-1}v{_{(1)}}){\otimes} v{_{(0)}}),
\eqno{(5)}
$$
where $\tau(a) = {\tau^{(1)}}}\def\taut{{\tau^{(2)}}(a){\otimes}{}_ B\taut(a)$ is a translation map in
$P(B,H)$, and then use properties of a translation map to prove that $\phi$ has
the
required properties and that the correspondence $\theta: \phi\mapsto s$ is
bijective.
{\ $\Box$}\bigskip
\remar{Example\ {3.2.2.}\ }
{ Let $E(B,V,H)$ be a quantum fibre bundle associated to a trivial
quantum principal bundle $P(B,H,\Phi)$ as described in
Example~2.3.2. In this case every element of
$E$ has the from $\sum_{i\in I}b_ i\Phi_ E(v_ i)$ for some
$b_ i\in B$ and $v_ i \in V$, and the bijection $\theta$ of the
proof of Theorem~3.2.1 reads
$$
\theta(\phi)(\sum_{i\in I}b_ i\Phi_ E(v_ i)) = \sum_{i\in
I}b_ i\Phi({\rm S}^{-1}v_ i{_{(1)}})\phi(v_ i{_{(0)}}),
$$
for any equivariant $\phi:V\to P$. The inverse of $\theta$ associates
an equivariant map $\theta^{-1}(s):V\to P$,
$$
\theta^{-1}(s)(v) = \Phi^{-1}({\rm S}^{-1}v{_{(1)}})s(\Phi_ E(v{_{(0)}}))
$$
to any $s\in\Gamma(E)$. Notice that the map $\theta^{-1}(s)$
obtained in this way is different from the equivariant map $\phi$
discussed in [6, Proposition A6].}
{_{(3)}}{Corollary }{3.2.3} {
Cross sections $s:P\to B$ of a quantum principal bundle $P(B,H)$ are in
bijective correspondence
with the maps $\phi :H\to P$ such that $\Delta_ R\phi =
(\phi{\otimes}{\rm S})\Delta'$ and $\phi(1) = 1$.}
Note that in Corollary~3.2.3. we do not
need the invertibility of ${\rm S}$, but
if $H$ has a bijective antipode ${\rm S}$, the
sections of
a quantum principal bundle $P(B,H)$ are in one-to-one correspondence
with the maps $\psi:H\to P$ such that $\psi(1) = 1$ and $\Delta_
R\circ\psi = ({\rm id}{\otimes}\psi)\circ\Delta$. We simply need to define
$\psi=\phi\circ{\rm S}^{-1}$, where $\phi$ is given by
Corollary~3.2.3.
{_{(3)}}{Proposition }{3.2.4 }{
Any trivial quantum principal bundle $P(B,H,\Phi)$ admits a section.
Conversely, if a bundle $P(B,H)$ admits a section which is an algebra
map then $P(B,H)$ is trivial with the total space $P$ isomorphic to
$B{\otimes} H$ as an algebra.}
\Proof A convolution inverse of a trivialisation $\Phi$ of a trivial quantum
principal bundle
$P(B,H,\Phi)$ satisfies the assumptions of
Corollary~3.2.3, hence $s={\rm id}*\Phi^{-1}$ is a
section of $P(B,H,\Phi)$. Conversely, assume that an algebra map
$s:P\to B$ is a section of $P(B,H)$. Clearly, $s$ is a $B$-bimodule
map, hence we can define a linear map $\Phi:H\to P$, $\Phi
=m\circ(s\otimes{}_ B{\rm id})\circ\tau$. One then shows that $\Phi$ is a
trivialisation and
$\tilde{\theta}(s)$ constructed in
Corollary~3.2.3 is its convolution inverse. {\ $\Box$}\bigskip
\remar{Remark\ {3.2.5.}\ }
{ We would like to emphasise that the existence of a cross section of a
quantum principal bundle does not necessarily imply that the bundle
is trivial. As an example of a non-trivial quantum principal bundle
admitting a cross section we consider the quantum Hopf fibration of
Example~2.2.4. We consider a linear map
$\phi : k[Z,Z^{-1}]\to SU_ q(2)$, given by
$$
\phi(1) = 1,\qquad \phi(Z^ n) = \delta^ n, \qquad \phi(Z^{-n})
=\alpha^ n,
$$
for any positive integer $n$. The map $\phi$ satisfies the hypothesis
of Corollary~3.2.3, hence it induces a cross section $s:
SU_ q(2) \to S_ q^ 2$, $s: u\mapsto u{_{(1)}}\phi(\pi(u{_{(2)}}))$ but
$s$ is not an algebra map since, for example,
$
s(\alpha\beta) = b_ -\neq q^{-1}b_ 3 b_ - = s(\alpha)s(\beta).
$}
\vskip4pt plus2pt
{\bf 3.3.} {\it Vertical automorphisms = gauge transformations}
\defin{Definition}{ 3.3.1.
Let $P(B,H)$ be a quantum principal bundle. Any left $B$-module
automorphism ${\cal F} : P\to P$ such that ${\cal F} (1) =1$ and $\Delta_ R
{\cal F} = ({\cal F}\otimes {\rm id})\Delta_ R$ is called a {\it vertical
automorphism} of the
bundle $P(B,H)$. The set of all vertical automorphisms of $P(B,H)$ is
denoted by $Aut_ B(P)$.}
Elements of $Aut_ B(P)$ preserve both the base space $B$ and the
action of the structure quantum group $H$ of a quantum principal
bundle $P(B,H)$. $ Aut_ B(P)$ can be equipped with a multiplicative
group structure $\cdot :({\cal F}_ 1, {\cal F}_ 2)\mapsto {\cal F}_
2\circ{\cal F}_ 1$. Vertical automorphisms are often called gauge
transformations and $ Aut_ B(P)$ is termed a gauge group.
{_{(3)}}{Proposition} {3.3.2.} {
Vertical automorphisms
of a quantum principal bundle $P(B,H)$ are in bijective correspondence
with convolution invertible
maps $f:H\to P$ such that $f(1)=1$ and $\Delta_ Rf = (f\otimes
{\rm id}){\rm Ad_{R}}$.}
\Proof If $f$ is a map satisfying the hypothesis of the proposition. then
${\cal F} ={\rm id}*f$. Conversely, for any ${\cal F}\in Aut_ B(P)$ a map
$f: A\to P$,
$f =m\circ({\rm id}{\otimes}{}_ B{\cal F})\circ\tau$ ,
where $\tau$ is a translation map has all the required properties. {\ $\Box$}\bigskip
Maps $f:H\to P$ form a
group with respect to the convolution product. This group is
denoted by ${\cal H}(P)$. There is an action of ${\cal H}(P)$ on
the space of connection one-forms in $P(B,A)$ given by
$(\omega, f)\mapsto \omega^ f = f^{-1}*\omega*f + f^{-1}*{\rm d} f$.
The connection one-form $\omega^ f$ is called a gauge transformation
of $\omega$. If $\omega$ is strong so is its gauge transformation.
Gauge transformation of such $\omega$ induces the gauge
transformation of its curvature $F\mapsto f^{-1}*F*f$.
Similarly there is an action of ${\cal H}(P)$ on $\Gamma(E)$ viewed
as equivariant maps $\phi: V\to P$ by Theorem~3.2.1.,
given by $(\phi, f)\mapsto\phi^ f =\phi*f$. These are the transformation
properties of the fields in quantum group gauge theories.
Proposition~3.3.2. implies the following:
{_{(3)}}{Corollary }{ 3.3.3. } {
For a quantum principal bundle $P(B,H)$, $Aut_ B(P)\cong{\cal H}(P)$ as
multiplicative groups.}
{_{(3)}}{Theorem }{ 3.3.4. }{
Let $P(B,H,\Phi)$ be a trivial quantum principal bundle. Then the
groups $Aut_ B(P)$, ${\cal H}(P)$, and the gauge group
${\cal H}(B)$ are isomorphic to each other.}
Therefore
Theorem~3.3.4. allows one to interpret a vertical
automorphism of a (locally) trivial quantum principal bundle as
a change of local variables and truly as a gauge transformation of a
trivial quantum principal bundle.
\section{4. Conclusions and open problems}
In this paper we reviewed basic properties
of quantum fibre bundles introduced in [6]. There is
a number of constructions, already present in the literature,
that we have not described in here.
For example, locally trivial quantum principal bundles,
defined in [6] were developed
by M. Pflaum in [23], using the methods of the sheaf
theory. A very interesting example of the Yang-Mills theory
in quantum bundles was constructed by P. Hajac in [17].
The example considered in [17]
belongs to the interface of the theory
described here and the Connes-Rieffel Yang-Mills
theory [11], and points to the very important problem of finding
the relationship between the quantum group gauge theory
and Connes' non-commutative geometry [9].
There is also a number of challenging problems
that need to be solved in order to obtain a full understanding
of quantum group gauge theories. For example, in this article
we restricted our discussion only to gauge transformations
of bundles with the universal differential structure. The theory
of gauge transformations of bundles with general differential
structures is not yet known. In particular, we would like to
define gauge transformations in such a way that a gauge
transformation of a connection one-form is still a connection
one-form. A couple of remarks on this problem may be found in
[3]. Also, it would be interesting to equip
our algebraic constructions with a some kind of topology,
like $C^ *$ or Frechet topology. Some topological aspects
of quantum fibre bundles are discussed in [8].
Furthermore, the theory of quantum
fibre bundles reviewed in this article is strictly related to
the theory of algebraic extensions. We think that the analysis
of quantum bundles from the point of view of Hopf-Galois
extensions may lead to a deeper insight into the both subjects.
Finally, we think it is desirable to develop generalised
fibre bundles defined in Remark~2.2.5. in order to construct
a gauge theory on general homogeneous spaces. The developement
of such a theory becomes even more important and challenging now
that the appearence of the $SU_q(2)$ homogeneous spaces
in the Connes description of Standard Model was announced [10].
\section{Acknowledgements}
Most of the results presented in this paper were
obtained jointly with Shahn Majid. I would like to thank him
for a fruitful collaboration and many interesting
discussions. This paper was written during my
stay at the Universite Libre de Bruxelles; I am grateful to the European Union
for the fellowship in the framework
of the Human Capital and Mobility Scheme. My work was also
supported by the grant KBN 2 P 302 21706 p 01.
\references{No}{\item{[1]}
R.J. \spa{Blattner,} M. \spa{Cohen} and S. \spa{Montgomery},
{\it Crossed Products and
Inner Actions of Hopf Algebras}\/, Trans. Amer.
Math. Soc. 298 (1986) 671.
\item{[2]} T. \spa{Brzezi{\'n}ski}, {\it
Differential
Geometry of Quantum Groups and Quantum Fibre Bundles}\/,
University of Cambridge, PhD thesis, 1994.
\item{[3]} T. \spa{Brzezi{\'n}ski}, {\it Remarks on Quantum Principal
Bundles}\/, In. {\it Quantum Groups. Formalism and
Applications}, J. Lukierski, Z. Popowicz and J. Sobczyk, eds.
Polish Scientific Publishers PWN, 1995, p. 3.
\item{[4]} T. \spa{Brzezi{\'n}ski}, {\it Translation Map in Quantum
Principal Bundles}\/, preprint (1994) {\tt hep-th/9407145}.
\item{[5]} T. \spa{Brzezi{\'n}ski}, {\it Quantum Homogeneous
Spaces as Quantum Quotient Spaces}\/, in preparation (1995)
\item{[6]} T. \spa{Brzezi{\'n}ski}
and S. \spa{Majid}, {\it Quantum Group Gauge Theory on Quantum Spaces}\/,
Commun.~Math.~Phys. 157 (1993) 591; {it ibid.}
167 (1995) 235 (erratum).
\item{[7]} T. \spa{Brzezi{\'n}ski}
and S. \spa{Majid}, {\it Quantum Group Gauge Theory on Classical
Spaces}\/,
Phys. Lett. B298 (1993) 339.
\item{[8]} R.J. \spa{Budzy{\'n}ski} and
W. \spa{Kondracki}, {\it
Quantum principal fiber bundles: topological aspects}\/,
preprint (1994) {\tt hep-th/9401019}.
\item{[9]} A. \spa{Connes}, {\it Non-Commutative Geometry}\/,
Academic Press, 1994.
\item{[10]} A. \spa{Connes}, {\it A lecture given at the
Conference on Non-commutative Geometry and Its Applications}\/,
Castle T\v re\v s\v t, Czech Republic, May 1995.
\item{[11]} A. \spa{Connes} and M. \spa{Rieffel}, {\it
Yang-Mills for Non-Commutative
Two-Tori}\/, Contemp. Math. 62 (1987) 237.
\item{[12]}C.-S. \spa{Chu},
P.-M. \spa{Ho} and H. \spa{Steinacker}, {\it
Q-deformed {D}irac monopole
with arbitrary charge}\/, preprint (1994) {\tt hep-th/9404023}.
\item{[13]} Y. \spa{Doi}, {\it Equivalent Crossed Products
for a Hopf Algebra}\/,
Commun. Algebra 17 (1989) 3053.
\item{[14]} Y. \spa{Doi} and M. \spa{Takeuchi}, {\it
Cleft Module Algebras
and Hopf Modules}\/, Commun. Algebra 14 (1986) 801.
\item{[15]} V.G. \spa{Drinfeld}, {\it Quantum Groups}\/, In {\it
Proceedings of the
International Congress of
Mathematicians, Berkeley, Cal. Vol.1}, Academic Press, 1986, p.798.
\item{[16]} T. \spa{Eguchi,} P. \spa{Gilkey} and A. \spa{Hanson},
{\it Gravitation,
Gauge Thoeries and Differential Geometry}\/, Phys. Rep.
66 (1980) 213.
\item{[17]} P.M. \spa{Hajac}, {\it Strong Connections
and $U\sb
q(2)$-Yang-Mills Theory on Quantum Principal Bundles}\/,
preprint(1994) {\tt hep-th/9406129}.
\item{[18]} D. \spa{Husemoller}, {\it Fibre
Bundles}\/,
Springer-Verlag, 3rd ed. 1994.
\item{[19]} D. \spa{Kastler}, {\it Cyclic Cohomology
within Differential Envelope}\/, Hermann, 1988.
\item{[20]}E. \spa{Kunz},
{\it K\"ahler Differentials}\/, Vieweg \& Sohn, 1986.
\item{[21]} S. \spa{Majid}, {\it Cross Product Quantisation,
Nonabelian
Cohomology and Twisting of Hopf Algebras}\/,
In. {\it Generalised Symmetries in Physics}, H.-D. Doebner,
V.K. Dobrev and A.G. Ushveridze, eds., World Scientific, 1994, p. 13.
\item{[22]} U. \spa{Meyer}, {\it Projective Quantum Spaces}\/,
Lett. Math.
Phys. to appear.
\item{[23]}
M. \spa{Pflaum}, {\it Quantum Groups on Fibre Bundles}\/,
Commun. Math.
Phys. 166 (1994) 279.
\item{[24]} P. \spa{Podle{\'s}}, {\it Quantum Spheres}\/, Lett. Math.
Phys. 14 (1087) 193.
\item{[25]} H.-J. \spa{Schneider} {\it Principal Homogeneous Spaces
for Arbitrary Hopf Algebras}\/, Israel J. Math. 72
(1990) 167; R.J. \spa{Blattner} and S. \spa{Montgomery}, {\it
Crossed Products and Galois
Extensions of Hopf Algebras}\/, Pacific J. Math. 137
(1989) 37.
\item{[26]} M.E. \spa{Sweedler}, {\it Hopf Algebras}\/,
Benjamin, 1969.
\item{[27]} M.E. \spa{Sweedler}, {\it Cohomology of Algebras
over Hopf
Algebras}\/, Trans. AMS 133 (1968) 205.
\item{[28]} S.L. \spa{Woronowicz}, {\it Twisted $SU\sb 2$ Group. An
Example of a Non-commutative Differential Calculus}\/,
Publ. RIMS Kyoto University 23 (1987) 117.
\item{[29]} S.L. \spa{Woronowicz}, {\it Differential Calculus on Compact
Matrix Pseudogroups (Quantum Groups)}\/,
Commun.~Math.~Phys.
122 (1989) 125.
}
\bye
|
\section{Introduction}
Discretized differential equations lie at the heart of many simulation
algorithms in physics. A large variety of solution algorithms like
Conjugate Gradient, Overrelaxation, or Multigrid exist to deal
efficiently with such problems \cite{Press}. The convergence of these
algorithms
usually depends on the condition number of the problem operator,
i.e.~the quotient of its largest and smallest eigenvalue. (For many
simple problems multigrid methods will always converge well. Here we
are not interested in such cases.) When the number of eigenmodes with
very small eigenvalues is not large, each of these methods could be
accelerated if an additional method for dealing with these modes would
be applied.
In this paper we want to study a
method that can be used to do exactly this. It is partly based on the
multigrid idea and relies on a surprisingly simple principle,
called the {\em Principle of indirect elimination\/} or {\em PIE}.
We will explain this principle in general context and then apply it to
a case where the occurence of almost-zero modes spoils the convergence
of standard methods, namely the Dirac equation in a gauge field
background with instantons \cite{Dilger,Sorin}. We will also show
connections to an idea by Kalkreuter somewhat similar in spirit,
called the {\em updating on the last point} \cite{ThomasII}, and
explain why our method is more general. Finally, we will briefly
remark on the connections to the Iteratively Smoothing Unigrid
algorithm \cite{ISU}.
\section{The general problem}
Consider a linear operator~${\bf D}$ which may arise from a discretized
differential equation. Here and in the following we assume~${\bf D}$ to
be {\em positive definite}, if it were not, we could use the
operator~${\bf D}^\ast {\bf D}$ instead. The general form of the equation to
be solved is then
\begin{equation}
\label{fundamental_eq}
{\bf D} {\xi} = {\bf f} \quad .
\end{equation}
Let us call the lowest eigenvalue of the operator~$\varepsilon_0$
\footnote {It is not fully correct to speak about eigenvalues
of~${\bf D}$. In section~\ref{Dual} we will explain what is meant
by such a statement.}. Its value determines the {\em criticality\/}
of the operator because the smaller it is the larger the condition
number (quotient of largest and smallest eigenvalue)
of the operator will be. If $\varepsilon_0=0$ the problem is
ill-posed because the contribution of this zero-mode to the solution
is not determined. For small $\varepsilon_0$ standard iterative methods
will converge only slowly,
the
convergence time~$\tau$ (the number of iterations needed to reduce the
error by a factor of~e) behaving like $\tau\propto \kappa^{z/2}$,
where $\kappa$ is the condition number of ${\bf D}$ and $z$ is the
critical exponent. This behaviour is called {\em critical slowing down\/}
because the more critical the problem gets the slower the algorithm
will be. For relaxation methods, one usually finds
$z\approx2$, Conjugate Gradient has a critical exponent of
$z\approx1$. An optimal algorithm should have a critical exponent of~0.
At each time-step, any iterative method will yield an approximate
solution~${\tilde{{\bxi}}}$.
We introduce two important quantities: the {\em error\/}~${\bf e}= {\xi}
-{\tilde{{\bxi}}}$ which is the difference between the true and the actual
solution and is of course not known, and the {\em residual\/}~${\bf r}= {\bf f}
- {\bf D} {\tilde{{\bxi}}}$, the difference between the true and the actual
righthandside. With these definitions we can recast the fundamental
equation~(\ref{fundamental_eq}) as
\begin{equation} {\bf D} {\bf e} = {\bf r} \quad, \label{error_eq} \end{equation}
called the {\em error equation}.
For a linear method, we can also introduce the {\em iteration matrix}~${\bf S}$
which tells us what the new error after the next iteration step will
be, given the old one:
\begin{equation} {\bf e}^{\rm new} = \matr{S} {\bf e}^{\rm old}
\quad. \label{IterError} \end{equation} The concrete structure of
the iteration matrix is irrelevant for the following discussion, see
\cite{Varga,Young} for examples. The important point here is that
the iteration matrix is the reduction matrix for the error. Its
eigenvalues should lie between minus one and one and convergence is
governed by the eigenmode of~${\bf S}$ with absolute value of the
eigenvalue closest to one.
In the following sections we will usually assume the algorithm to be
linear because the existence of an iteration matrix eases the
analysis. Nevertheless the method presented here could be applied to
the Conjugate Gradient algorithm as well, see also section~\ref{instantons}
\subsection{Remark on vector spaces}
\label{Dual}
For the analysis it is important to distinguish between a vector space
and its dual \cite{Sokal}. The differential operator~${\bf D}$ maps a
vector ${\xi}\in V$ to a vector in the dual space ${\bf f} \in V^*$.
To see this, consider the Laplace equation in electrodynamics as an
example: $\Delta \varphi = - \varrho$. The Laplace operator maps a
potential onto a charge density. These two objects can be regarded as
dual vectors because there is a unique way of assigning a real number
to them, namely the energy $\int \varrho(x) \varphi(x) dx$. The Laplace
operator therefore provides us with a bilinear form
$\bracket{\varphi}{\psi}_\Delta = \int \varphi(x) (\Delta \psi)(x)
dx$. However, there is no natural identification between the vector
space and its dual besides that given by this scalar product.
We will later see an example where one is easily drawn to wrong
conclusions if this distinction is not taken into account.
It is not really meaningful to speak about eigenvectors or -values of
bilinear forms. On the other hand, the iteration matrix of
relaxational methods maps the error to another error and is therefore
a map ${\bf S} : V \rightarrow V$, possessing eigenvectors. It are the
eigenvalues of this matrix that determine the convergence. The
standard identification of eigenvectors of ${\bf S}$ with eigenvectors of
${\bf D}$ is done using additional structure. This is given by the matrix
${\bf B}_0$ which is defined through the relation ${\bf S} = {\bf I} - {\bf B}_0^{-1}
{\bf D}$. (Standard relaxation methods arise from splitting the
fundamental operator ${\bf D} = {\bf B}_0 + {\bf C}_0$, where ${\bf B}_0$ is chosen such
that it approximates ${\bf D}$ as good as possible but is ``easy to
invert''.) ${\bf B}_0$ is an additional bilinear form and furnishes us with
a scalar product in addition to the scalar product given by ${\bf D}$.
For Conjugate Gradient, the situation is similar: Conjugate Gradient
updating steps require computations of scalar products, e.g.\ $\alpha
= \bracket{{\bf r}}{{\bf r}} / \bracket{\bf d}{{\bf D}{\bf d}}$, where ${\bf d}$
is the search vector. Here we need another scalar product than the
${\bf D}$-product.
It is therefore only correct to speak of eigenvectors of ${\bf D}$ when we
have chosen a basis that is in some sense natural. For example, if we
use the standard site-wise basis and find that the eigenvectors of
${\bf D}$ in this basis agree with those of ${\bf S}$, the sloppy way of speach
is justified. This will be the case for the example we will
study below. Nevertheless, in the theoretical parts of this paper
we will be more strict.
\section{PIE in general}
After these preliminaries we formulate the
\begin{quote}
{\bf Principle of indirect elimination (PIE):\/} It is easier to {calculate\/}
the shape of a bad-converging mode for a certain algorithm than to
reduce it directly using this algorithm.
\end{quote}
To see this, consider the case where there is only one bad-converging
mode and all others are reduced efficiently by the algorithm. We now
use the algorithm to
{\em try\/} to solve an equation of which we already know the
solution, for example
the equation ${\bf D} {\xi} =0$. In this case we have ${\tilde{{\bxi}}} = -{\bf e}$,
so we know the error as well. Remembering equation~(\ref{IterError})
we see that we can now directly investigate how the iteration matrix
acts. After~$n$ iterations we have
\begin{equation}{\tilde{{\bxi}}}^{(n)} = {\bf S}^n {\tilde{{\bxi}}}^{(0)} \quad , \end{equation}
where ${\tilde{{\bxi}}}^{(0)}$ is the initial guess we started with and
${\tilde{{\bxi}}}^{(n)}$ is the approximate solution after the $n$-th
iteration. For $n\rightarrow\infty$ ${\bf S}^n$ projects onto the
eigenvector of ${\bf S}$ with the largest absolute value of the
eigenvalue, which is the slowest-converging mode. For
finite $n$ the accuracy of the projection depends on quotient
between the largest and the second-largest eigenvalue: The larger
this is, the better the projection will be. (This can be seen easily
by imagining ${\bf S}$ to be diagonalized.) In the model case considered
here, where there is only one bad-converging mode, this quotient
will be large and so ${\bf S}^n {\tilde{{\bxi}}}^{(0)}$ will converge rapidly
against the bad-converging mode.
If the number of bad-converging modes is larger than one, but still
small, we can use the same technique to calculate them if we take
care of orthogonalizing the approximations to the already known
modes. By this it is obvious that this method will only be useful if
this number is not too large, otherwise the calculations will take too
much time. We will later comment on how the principle of indirect
elimination can be applied locally and used to construct a multigrid
algorithm.
Let us come back to the case of only one bad-converging mode. If we
have calculated this using the principle of indirect elimination, how
can we apply this knowledge to improve convergence?
The answer relies on multigrid ideas and is in fact very simple. Let
us call the bad-converging mode ${\bf w}$. We define an operator $\calA:
{\bf R}\rightarrow V, \mu \mapsto \mu {\bf w}$ that creates a vector on the
fundamental lattice from a number. This cumbersome notation has a
two-fold purpose: First it stresses the similarity to multigrid ideas,
where $\calA$ would be called an interpolation operator, second it will
later allow us to study the case where $\calA$ is not exactly equal to
the bad-converging mode ${\bf w}$ to see how this will affect the
convergence.
To solve the inhomogenuous
equation, we first apply our standard iterative solver a few times.
This will reduce all components of the error appreciably except for a
part proportional to~${\bf w}$: ${\bf e} \approx c {\bf w}$.
Inserting this knowledge into the error equation~(\ref{error_eq}) or
using the fact that ${\bf r} = {\bf D} {\bf e} \approx {\bf D}{\bf w} c$ we get
\begin{equation}
{\bf D} (c \calA) \approx {\bf r} \Longrightarrow \calA^* {\bf D} \calA c \approx \calA^* {\bf r} \quad .
\end{equation}
In other words, we have transformed the fundamental equation, living
on a large lattice, into an equation for scalars (or simple matrices
in the case of a gauge theory, see below). This new equation can be
considered to live on a lattice with only {\em one point}. In
multigrid language this is often called the ``last-point lattice'' as
we have there a whole tower of coarser and coarser lattices of which
the last consists of only one point.
The equation on the last point can be solved easily to get~$c$ and
afterwards we correct our approximation: ${\tilde{{\bxi}}} \leftarrow
{\tilde{{\bxi}}} + \calA c$. Thus we have reduced that part of the error
corresponding to~$\calA$. It is well-known from the multigrid context
that using the largest mode of~${\bf S}$ as interpolation operator will
yield the best convergence (Greenbaum criterion \cite{Greenbaum}). If
the iterative method used before has not been perfect, i.e.~if the
error still contains contributions from other modes, we now have to
start the iteration again to act on the remaining parts. This may
again introduce error-components proportional to $\calA$ which are then
reduced by another ``last-point updating''.
We can now understand the reason why the principle has been called
principle of indirect elimination: Direct elimination of the
bad-converging mode using the iterative solver does not work
efficiently, but an indirect approach, first trying to solve an
auxilliary equation and only afterwards addressing the real problem,
works fine.
In practice the situation will not be the idealized one described
above. We now want to study two situations: What will the result of
the correction be when the error is not an exact
multiple of the zero-mode~${\bf w}$, and what happens when $\calA$ deviates
{} from ${\bf w}$?
In the first case it is easy to prove that
after the
correction ${\tilde{{\bxi}}}$ and $\calA$ will be ${\bf D}$-orthogonal even if the
error before was not a multiple of $\calA$, but contained an additional
contribution ${\bf v}$:
\[
-{\tilde{{\bxi}}} = {\bf e} = c \calA + {\bf v} \]
\[ {\bf r} = {\bf D} {\bf e} = {\bf D} \calA c + {\bf D} {\bf v}\]
{\hbox{The equation on the last point is then:}}
\[ \calA^* {\bf D} \calA\, x = \calA^* {\bf D} \calA\, c + \calA^* {\bf D} {\bf v}\]
\[ \Longrightarrow \qquad x = c +
\frac{\displaystyle\calA^* {\bf D} {\bf v}}{\displaystyle\calA^* {\bf D} \calA}\]
{\hbox{Correcting ${\tilde{{\bxi}}}$ yields}}
\[ {\tilde{{\bxi}}}
= \frac{\displaystyle\calA^* {\bf D} {\bf v}}{\displaystyle\calA^* {\bf D} \calA}\, \calA
- {\bf v}\]
{\hbox{This gives ${\bf D}$-orthogonality: }}
\[ \calA^* {\bf D} {\tilde{{\bxi}}} = + \frac{\displaystyle\calA^* {\bf D}
{\bf v}}{\displaystyle\calA^* {\bf D} \calA}\, \calA^* {\bf D} \calA - \calA^* {\bf D} {\bf v} =
0 \]
Now we want to investigate the second question, namely how well the
approximation of the zero-mode has to be. To do so we can prove the
following rather trivial
\begin{thm}
We have an algorithm consisting of two parts. The
first part is able to eliminate completely all components of the
error except one single mode~${\bf w}$, so we have ${\bf e}={\bf w}$. The
second updating then consists of an updating on the last point as
described above using an approximation $\calA$ of ${\bf w}$.
We can split the bad-converging mode~${\bf w}$ into two ${\bf D}$-orthogonal
parts:
\begin{equation} {\bf w} = {\cal A} c + {\bf v} \quad, {\rm \ with\ }
\langle{\cal A} ,{\bf D}{\bf v}\rangle=0 \quad.\end{equation}
Then the iteration matrix ${\bf M}$ of the full algorithm consisting
of both steps has the (squared) energy norm (with respect to~${\bf D}$)
\begin{equation} \|{\bf M}\|_{\bf D}^2 =
\frac{\langle{\bf v},{\bf D}{\bf v}\rangle}{\langle{\bf w},{\bf D}{\bf w}\rangle}
\quad .\end{equation}
\end{thm}
\begin{pf}
The energy norm is defined as
\[ \|{\bf M}\|_{\bf D}^2 =
\sup_{\xi} \frac{\langle{\bf M}{\xi}, {\bf D}{\bf
M}{\xi}\rangle}{\langle{\xi},{\bf D}{\xi}\rangle} \quad.\]
Let ${\bf S}$ be the iteration matrix of the first part of the algorithm.
As it eliminates all parts of an arbitrary error except the
mode~${\bf w}$, it is clear that the supremum in the definition will be
reached for~${\xi}={\bf w}$. ${\bf S}$ does not affect~${\bf w}$.
After the iteration only that part of ${\bf w}$ that is ${\bf D}$-orthogonal
to $\calA$ will remain, see the calculation above.
So we get ${\bf M} {\bf w} = {\bf v}$.
Thus we have
\[ \|{\bf M}\|_{\bf D}^2 =
\frac{\langle{\bf M}{\bf w}, {\bf D}{\bf
M}{\bf w}\rangle}{\langle{\bf w},{\bf D}{\bf w}\rangle} =
\frac{\langle{\bf v},{\bf D}{\bf v}\rangle}{\langle{\bf w},{\bf D}{\bf w}\rangle}
\quad.\qquad\hfill\qed
\]
\end{pf}
It is also useful to look at this geometrically:
The
angle~$\theta$ between
the vector~${\bf w}$ and ${\cal A} c$ with respect to the scalar product
defined by~${\bf D}$ is given by
\[ \cos\theta =
\frac{\langle{\bf w},{\bf D}{\cal A} c\rangle}{\langle
{\bf w},{\bf D}{\bf w}\rangle^{1/2} \langle{\cal A} c
,{\bf D}{\cal A} c\rangle^{1/2}}\quad. \]
The
reduction works by first projecting ${\bf w}$ onto the direction given
by ${\cal A}$ and then taking the ${\bf D}$-orthogonal part of this. This
orthogonal part is the vector~${\bf v}$; it is all that remains after
the coarse-grid correction step. The length of this vector is given
by $\|{\bf v}\|_{\bf D} = \|{\bf w}\|_{\bf D} \sin \theta$. The reduction factor,
which is equal to the norm of the iteration matrix,
is~$\sin\theta$:
\begin{center}
\epsfig{file=lastpoint2.eps,width=5cm}
\end{center}
Using Pythagoras' theorem we get
\[
\sin^2\theta = 1-\cos^2 \theta = 1 -
\frac{\langle{\bf w},{\bf D}{\cal A} c\rangle^2}{\langle
{\bf w},{\bf D}{\bf w}\rangle \langle{\cal A} c
,{\bf D}{\cal A} c\rangle}\quad,\]
and inserting the split of the vector ${\bf w}$ and again using the
orthogonality property, we finally arrive at
\[
\sin^2\theta = 1 - \displaystyle\frac{ (\langle{\cal A} c,
{\bf D}{\cal A} c\rangle
+ \langle{\bf v},{\bf D}{\cal A} c\rangle)^2}{\langle
{\bf w},{\bf D}{\bf w}\rangle \langle{\cal A} c
,{\bf D}{\cal A} c\rangle}
= \displaystyle\frac{\langle{\bf w},{\bf D}{\bf w}\rangle-
\langle{\cal A} c,{\bf D}{\cal A} c\rangle}{
\langle{\bf w},{\bf D}{\bf w}\rangle}
= \displaystyle\frac{\langle{\bf v},{\bf D}{\bf v}\rangle}{\langle{\bf w},{\bf D}{\bf w}\rangle}
\quad.\]
What are the implications of this theorem? First it must be understood
that the energy norm of the iteration matrix will be equal to the
spectral radius provided the matrix~${\bf S}$ is ${\bf D}$-symmetric,
i.e.\ ${\bf S}^* {\bf D} = {\bf D} {\bf S}$ \footnote{Actually, this is a nice
example for the necesssity to distinguish endomorphisms and bilinear
forms: treating ${\bf D}$ as an endomorphism, not as a bilinear form, we
would transform it using the wrong relation and {\em loose\/} the
${\bf D}$-symmetry property after the transformation: We have ${\bf S}^* {\bf D}
= {\bf D} {\bf S}$. Now if we transform both using the transformation for
endomorphisms, we get $ ({\bf S}^\prime)^* {\bf D}^\prime = \left({\bf U}^* {\bf S}^*
({\bf U}^{-1})^*\right) {\bf U}^{-1} {\bf D} {\bf U}$! Here there is no cancellation as it
would be with the correct transformation: $ {\bf S}^{T^\prime} {\bf D}^\prime
= \left({\bf U}^* {\bf S}^* ({\bf U}^{-1})^*\right) {\bf U}^* {\bf D} {\bf U}$. Only if ${\bf U}$ is
orthogonal or
unitary do we get the same cancellation.}. This is true when the
operator~${\bf D}$ does not mix the mode~${\bf w}$ with the other modes. We
can then regard ${\bf w}$ as an eigenmode of ${\bf D}$ because the
matrix~${\bf S}$ provides us with an identification of the mode ${\bf w}$
with the corresponding mode in the dual space. Hence the energy norm
directly tells us about the convergence rate of the algorithm. ${\bf S}$
will also be ${\bf D}$-symmetric for standard iterative methods
like Jacobi-relaxation and can be made so as well for SOR or
Gau{\ss}-Seidel relaxation.
By choosing~${\bf w}$ as the zero-mode the theorem shows how important
the correct treatment of this mode is. The closer the range of the
interpolation operator ${\cal A}$ is to the zero-mode and the smaller the
energy norm of the residual part ${\bf v}$ the better the convergence
will be. (In the limiting case where the two are identical, the
difference vector is zero and the error of the zero-mode is eliminated
perfectly, as expected.) It is not only important that the
zero-mode is approximated well by the interpolation operator on the
last point, the convergence will also be better when the
difference vector between zero-mode and the mode used for the
interpolation has a small energy norm and is as smooth as possible.
The theorem also serves to explain a finding by Kalkreuter
\cite{ThomasII}. He found that it is possible to eliminate critical
slowing down in a multigrid algorithm (actually a two-grid) for the
standard Laplace equation even with interpolation operators that are
not able to represent the zero-mode (which is a constant in this case)
exactly, but only approximately. In the light of our theorem this
could be understood if the difference vector has a small energy norm.
This, however, has not been tested.
Thus we have seen the importance of the correct treatment of the
zero-mode. Other methods to remove convergence problems caused by the
zero-mode can be thought of. Kalkreuter \cite{ThomasII} proposed a
simple rescaling of the approximate solution~${\tilde{{\bxi}}}$ in addition
to a multigrid or a relaxation algorithm to improve the convergence.
This method completely eliminates critical slowing down in the
simplest model problem, the Laplace equation on a two-point grid.
The rescaling amounts to using the approximate solution ${\tilde{{\bxi}}}$
itself as interpolation operator~${\cal A}$.
The motivation for this updating scheme can be found in the following
argument: Consider again the equation
${\bf D} {\xi} =
{\bf f}$. Solving this gives ${\xi} = {\bf D}^{-1} \vect{f} = ({\bf B}_0^{-1}
{\bf D})^{-1} {\bf B}_0^{-1}f $ where we have inserted a unit matrix.
Let us fix the righthandside and increase the criticality of the problem.
The more critical it gets the smaller the lowest eigenvalue
$\varepsilon_0$ of $({\bf B}_0^{-1} {\bf D})$ will be.
The solution ${\xi}$ will then have larger and larger contributions
{} from the lowest eigenmode of $({\bf B}_0^{-1} {\bf D})$. Therefore ${\xi}$ itself
will be a good approximation to the bad-converging mode and can be
used as interpolation operator.
We see that this method is very similar to
the principle of indirect elimination. However, the principle of
indirect elimination will always provide us with an approximation to
the zero-mode without any contribution from a given righthandside,
whereas Kalkreuter's method will only work well for large
criticality: The interpolation operators used by the methods are $\calA =
{\bf S}^n {\xi}^{(0)}$ for the principle of indirect elimination and $\calA = {\bf S}^n
{\xi}^{(0)} + {\bf B}_0^{-n} {\bf f}$ for Kalkreuter's method.
Even more important, the principle of indirect
elimination can be used several times to remove more than just one
mode, this is impossible with the other approach. On the other hand for
large criticality and the case of only one bad-converging mode,
Kalkreuter's method has the advantage of not needing auxilliary
iterations to calculate $\calA$ because ${\tilde{{\bxi}}}$ is used.
Kalkreuter used this method in addition to a usual multigrid
or relaxation method. He found that there is no strong improvement for
a multigrid algorithm, but for standard local relaxation the
asymptotical critical slowing down (i.e.~critical slowing down for
fixed grid size and infinitely many iterations) was eliminated for the
Laplace equation with periodic boundary conditions. This is what we
expect for a method that treats the lowest mode of the problem correctly
because in this case it is the eigenvalue of the second-lowest mode
that determines the convergence and this scales with the size of the
grid, not with the lowest eigenvalue. So for increasing grid
sizes critical slowing down should still be present; this in agreement
with Kalkreuter's results.
\section{Killing Instantons}
\subsection{The Dirac operator}
Our example for a discretized differential equation with a small
number of bad-converging modes is taken from theoretical high-energy
physics, namely the two-dimensional Dirac equation on the lattice in a
gauge field background with periodic boundary conditions.
For an introduction to Lattice Gauge Theory consult \cite{Creutz}.
Here we only present the framework:
Consider a regular, $d$-dimensional (hyper-)cubic lattice $\Lambda^0$
with lattice constant~$a$, lattice points~$z$ and directed
links~$(z,\mu)$. The opposite link is then denoted by $(z+\mu,-\mu)$,
where $z+\mu$ means the next neighbour of $z$ in $\mu$-direction. The
direction index~$\mu$ runs from $-d$ to $d$. Usually, the lattices
used will be finite with an extension of $L$ points in each dimension
so that the number of degrees of freedom is~$n=L^d$.
A lattice gauge theory is defined by a gauge group~$G$ which might for
example be U(1) or SU(2). Elements of the gauge group act on a vector
space~$V$ which for the examples above would be ${\bf C}$ and ${\bf C}^2$,
respectively. The computations presented below were done in two
dimensions with gauge group U(1), so that instantons can occur. A
lattice gauge field (in this case) assigns a U(1)-``matrix''
$U(z,\mu)$ (which is simply a phase) to every link of the lattice,
subject to the condition $U(z,\mu) = U(z+\mu,-\mu)^{-1}$. These
matrices are distributed randomly with a Boltzmannian probability
distribution $\propto \exp (-\beta S_W(U))$, where $S_W$ is the
standard Wilson action of lattice gauge theory
\[S_W(U)= \sum_p Tr (1-U(\partial p)) \quad {\rm with}\ U(\partial p)
= U(b_4)U(b_3)U(b_2)U(b_1)\] for a plaquette~$p$ of the lattice with
links $b_1... b_4$ at its boundary. This distribution leads to a
correlation between the gauge field matrices with finite correlation
length $\chi$ for finite $\beta$. The case $\beta = 0$ corresponds to
a completely random choice of the matrices ($\chi=0$), for
$\beta=\infty$ all matrices are $\matr{1}$ ($\chi=\infty$). In this
sense, $\beta$ is a disorder parameter, the smaller $\beta$ the
shorter the correlation length and the larger the disorder.
The Dirac operator acts on matter fields~${\xi}$ living on the nodes of
the lattice. In Kogut-Susskind formulation \cite{Staggered} it is
defined as
\[
\left( {\,{\bf D}\!\!\!\! /\,} {\xi}\right)(z) =
\frac1a \sum_{\mu=1}^d \eta_{\mu, z} \left(U(z,\mu)^*\,\xi(z+\mu) -
U(z,-\mu)^* \,\xi(z-\mu) \right)
\quad.\]
Here the~$\eta_{\mu,z}= \pm 1$
are the remnants of the Dirac matrices~$\gamma^\mu$ in the continuum.
As the Dirac operator itself is not positive definite, we will use its
square~${\,{\bf D}\!\!\!\! /\,}^2$ in the following. The squared Dirac has the
property of totally decoupling the even and odd parts of the lattice;
if we color the lattice points in checkerboard fashion, any red point
is only coupled to other red points, so that we can restrict our
attention to one of the sub-lattices. This will be especially useful
because it lifts the degeneracy of the eigenvalues: Usually each
eigenvalue of the Dirac operator is degenerated twice, but we can
choose the eigenvectors to live separated on the sub-lattices. For
the sake of brevity we will generally speak of the Dirac operator even
when we
mean the squared Dirac.
\subsection{The instanton problem}
\label{instantons}
Many algorithms for solving the Dirac equation become problematic in
the presence of instantons. Instantons are gauge field configurations
that are topologically non-trivial but possess zero energy. Such
configurations are only possible for certain choices of the dimension
and the gauge group, in two dimensions instantons can occur when the
gauge group is U(1), see \cite{Nakahara} for an introduction.
The Atiyah-Singer theorem states that at
instanton charge~$Q$
the spectrum in the continuum will possess $2|Q|$ exact
zero-modes~\cite{Atiyah}; these become modes with extremely small
eigenvalues on the lattice \cite{Dilger}. For the purpose of this
paper it is not necessary to have an understanding of what an
instanton is, it is only important that they are special gauge field
configurations giving rise to almost-zero-modes on the lattice.
Figure~\ref{Qtopspectrum} shows the lower part of the spectrum of the
squared Dirac operator on an $18^2$-lattice at $\beta=10$ for
different instanton charges~$Q$, taking only one of the two
sublattices into account. Clearly the Atiyah-Singer theorem is
nicely reflected on the lattice.
\begin{figure} \begin{center}
\epsfig{file=qtopspectrum.eps,width=8cm}
\end{center}
\caption[The Atiyah-Singer theorem on the lattice]{The Atiyah-Singer
theorem on the lattice: Shown are the five
lowest eigenvalues of the staggered squared Dirac operator in a
U(1) gauge field at~$\beta=10$ on an~$18^2$-lattice for different
topological charges~$Q$. Only one of the two sub-lattices is taken
into account, on the full lattice each eigenvalue is degenerated twice.
\label{Qtopspectrum}}
\end{figure}
When instantons are present,
the condition number of the Dirac operator becomes very large and the
problem is ill-posed. In \cite{Sorin} this problem is investigated in
detail for the Parallel Transported Multigrid. In this section, we
want to use the principle of indirect elimination to show how an
algorithm which converges well in the absence of instantons can be
adapted to a case with instantons.
The idea is very simple: If~$m$ bad-converging modes are present, use
the principle of indirect elimination $m$ times to calculate
approximations $\calA^i$ to these modes.
The method presented here could be applied to a Conjugate Gradient
algorithm. For this algorithm, Dilger \cite{Dilger} has found
that the number of iterations needed strongly increasses with the
instanton charge, therefore Conjugate Gradient would benefit from
the application of the method described here.
However, we will choose the ISU algorithm on small lattices and at
quite large values of~$\beta$ for a U(1) gauge field as an example.
This algorithm has been described in detail in~\cite{ISU}. As it is in
some parts based on the principle of indirect elimination, some
remarks will be made on this method in the next section. For this
section it is not necessary to understand how ISU works, it suffices
to know that for the parameters chosen the ISU algorithm converges
well for instanton charges 0 or $\pm1$ at large $\beta$ but badly for
larger instanton charges. The reason is that the algorithm in its
standard form contains one interpolation operator on the last point
(which is calculated as an approximation to the zero-mode) and so it
is able to eliminate one zero-mode, but not more.
In the improved algorithm
one tries to solve the equation ${\bf D} \calA^i=0$ with the given algorithm.
As it eliminates all other modes quickly the approximate
solution will converge against a linear combination of the
bad-converging modes. Then we start the procedure again, but now
orthogonalizing the approximate solution to the interpolation operator
we already know, doing this successively for all~$m$ bad-converging
modes. (As the instanton charge can be easily measured, one usually
knows beforehand how many operators are needed; if one does not for
some reason, a dynamical approach can be chosen: Simply proceed
calculating the next interpolation operator until the convergence rate
of the trivial equation becomes good enough.)
The overall work for this procedure is proportional to the square of
the number of bad-converging modes, as is the work of actually
applying the interpolation operators to eliminate them. (The number
gets squared because of the need to calculate an effective
operator on the last point layer. However, the effective operator only
needs to be calculated once for each configuration.) This restricts
the method to cases where the number of bad-converging modes is not
too large, which usually is the case for instanton charges.
\begin{figure} \begin{center}
\epsfig{figure=inst_killing.eps,width=6cm}
\end{center}
\caption[Performance of standard and improved ISU for instanton
problems]{Performance of the standard and the improved ISU algorithm
for the elimination of instanton modes. The data were generated on
one sub-lattice of
an~$18^2$-lattice at $\beta=10$ with a U(1) gauge field. The
improved algorithm uses $Q$~interpolation operators on the last
point to eliminate the almost-zero modes. (The number of
configurations evaluated for the different topological charges
was~217, 113, and 22. )
\label{KillingPic}}
\end{figure}
Figure~\ref{KillingPic} shows the performance of the usual ISU method
compared to the improved version for the Dirac operator in a U(1)
gauge field with different instanton charges. We measured the
asymptotic convergence time, i.e.~the number of iterations
asymptotically needed to reduce the error by a factor of~e. The
improved version of ISU used a number of interpolation operators on
the last point equal to the instanton charge which equals the number
of bad-converging modes. The data were generated on one sub-lattice of
an $18^2$-lattice at $\beta=10$. For this high value of $\beta$, the
convergence in the absence of instantons is good, as can be seen from
the value at $Q=1$. The standard method works well for instanton
charge~0 or 1, as explained above, and its sensitivity to higher
instanton charges is striking. The improved method shows no
dependence on the instanton charge, the convergence is good in all
cases. Note also that the standard deviation is much higher for the
usual method because it is affected by fluctuations in the eigenvalues
of the bad-converging modes.
Clearly the improved method is superior---the cost of
calculating the instanton modes is about 10~iterations for each
instanton plus the cost of the orthogonalization, whereas the saving
in the solution of the final equations is of the order of hundreds of
iterations depending on how much we want to reduce the residual.
\section{PIE and ISU}
We have seen above that the principle of indirect elimination will
only be helpful when the number of bad-converging modes is not too
large. In the case of simple relaxation methods, however, this number
is of order ${\cal O}(n)$, so
storing them would cost ${\cal O}(n^2)$, where $n$ is
the number of degrees of freedom in the system. So it seems that the
method is useless in such cases.
In this section we want to explain how the Iteratively Smoothing
Unigrid algorithm (ISU) presented in
\cite{ISU} can be regarded as the local application of the principle
of indirect elimination. We will only present the basic idea here.
Some familiarity with the basic multigrid idea is assumed in this
section, see \cite{Brandt,Briggs,Hack} for introductions.
We can associate a length scale (e.g.\ a wavelength) with each modes
of our system. Because they are local, relaxation methods eliminate
all those modes corresponding to a length scale of the order of one
lattice spacing. Usually there will be ${\cal O}(n/2)$ of these. Of the
remaining modes ${\cal O}(n/4)$ will be associated with length scale $2a$
($a$ is the lattice spacing), ${\cal O}(n/8)$ with scale $4a$ and so on.
So there will be many bad-converging modes with a small length scale
and only a few corresponding to a large scale. The ISU algorithm is a
method to calculate interpolation operators that are able to span the
space of these modes. These operators are restricted to parts of the
lattice, the size of the domain being determined by the scale of the
mode that is to be approximated by the operator.
To be more specific, let us start with the smallest scale $2a$. As in
usual multigrid methods, we divide the hypercubic lattice into
(overlapping) hypercubes or blocks $[x]$ of side length~3. Then we try
to solve the equation ${\bf D}|_{[x]} {\cal A}_x^{[1]}(z) =0$ using a
relaxation method. Here ${\bf D}|_{[x]}$ is the restriction of ${\bf D}$ to
the block $[x]$ using Dirichlet boundary conditions. What remains
after a few iterations will be the slowest-converging mode on this
scale and can therefore be used as interpolation operator on the first
block-lattice. Repeating this for all the small hypercubes, we know
the shapes of the bad-converging modes on scale $2a$. Now we do the
same on the next scale, dividing the lattice into larger blocks (of
side length~7, agreeing with the formula $2^j-1$). Again we try to
solve the equation ${\bf D} |_{[x]} {\cal A}_x^{[2]}(z) = 0$, where $[x]$
now denotes the larger blocks. The important point is that we use the
interpolation operators on the smaller scale that are already known
for this calculation to eliminate contributions from the
bad-converging modes on the smaller scale. In this way we proceed to
larger and larger hypercubes, always using the interpolation operators
already known. This method would only fail if a large number of
bad-converging modes lived on a large length-scale.
It has been found that this algorithm is able to eliminate critical
slowing down completely for the case of the two-dimensional Laplace
equation in an SU(2) gauge field background at arbitrarily large
values of the gauge field disorder and the lattice size. An improved
version has been shown to do the same for the two-dimensional squared
Dirac equation, except for extremely large disorder ($\beta \approx
2$ or smaller). See \cite{ISU} for details.
An idea that is similar in spirit to the principle of indirect
elimination has been discussed in \cite[section~4.6]{Achi}. Brandt
proposes to do relaxations on the fundamental lattice with arbitrary
starting vectors to determine ``typical shapes of a slow-to-converge
error'' which could then be used to determine good multigrid
interpolation operators. Unfortunately, this idea suffers from a
severe disease: The number of modes that converge badly under simple
relaxation is huge (about half of the number of grid points). What one
will get by this procedure is a mixture of low-lying eigenmodes with
contributions depending on their eigenvalues. The time needed to
arrive at a function that consists only of the lowest eigenmodes will
be proportional to the lattice size, so the method will not work
without critical slowing down.
The difference to the ISU algorithm is that this is a so-called {\em
unigrid} method. It allows for interpolation operators living on
different length-scales, whereas standard multigrid algorithms only
use interpolation operators living on small domains. On each length
scale we need not represent all modes that converge badly on this and
on all higher scales; only the modes that belong to the scale
corresponding to a certain lattice constant have to be dealt with. The
next-coarser length-scale will then take care of the modes
corresponding to this scale and in their computation the smaller
scales are already taken into account.
\section{Conclusions}
We have presented a simple method to improve the convergence of
solution algorithms for discretized differential equations when the
number of bad-converging modes is small. The principle of indirect
elimination used to do this is based on the general idea that an
algorithm can be used to identify its own bad-converging modes.
Conceptionally, the method is similar to the general idea of
accelerating algorithms described in \cite{Sorin2}: One tries to find
out what the slow modes of the algorithm are and uses this knowledge
to improve the algorithm. For example, multigrid methods are based on
the fact that the slow modes are the smooth modes that can be obtained
by smooth interpolation. The principle of indirect elimination serves
to automatize this process in the case when the number of slow modes
is small so that it suffices to know them without doing further
analysis of their structure.
The
method has been studied for the case of the Dirac equation in a gauge
field background with instantons and worked extremely well. Applying
it locally leads to a multigrid method called the Iteratively
Smoothing
Unigrid.
\begin{ack}
I wish to thank Gerhard Mack and Alan Sokal for stimulating
discussions. Hermann Dilger provided me with a copy of his program for
generating U(1) gauge field configurations.
Financial support by the Deutsche Forschungsgemeinschaft
is gratefully acknowledged.
\end{ack}
|
\section{#1}}
\renewcommand{\theequation}{\Roman{section}..\arabic{equation}}
\def\bseq{\begin{subequation}}
\def\end{subequation}{\end{subequation}}
\def\bsea{\begin{subeqnarray}}
\def\end{subeqnarray}{\end{subeqnarray}}
\evensidemargin 0.4cm
\oddsidemargin 0.4cm
\textwidth 15cm
\textheight 8.5in
\topmargin -1.2cm
\headsep .4in
\newcommand{\begin{equation}}{\begin{equation}}
\newcommand{\end{equation}}{\end{equation}}
\newcommand{\begin{eqnarray}}{\begin{eqnarray}}
\newcommand{\end{eqnarray}}{\end{eqnarray}}
\newcommand {\non}{\nonumber}
\renewcommand{\a}{\alpha}
\renewcommand{\b}{\beta}
\renewcommand{\c}{\gamma}
\renewcommand{\d}{\delta}
\newcommand{\theta}{\theta}
\newcommand{\Theta}{\Theta}
\newcommand{\partial}{\partial}
\newcommand{\partial}{\partial}
\newcommand{\gamma}{\gamma}
\newcommand{\Gamma}{\Gamma}
\newcommand{\Alpha}{\Alpha}
\newcommand{\Beta}{\Beta}
\newcommand{\Delta}{\Delta}
\newcommand{\epsilon}{\epsilon}
\newcommand{\zeta}{\zeta}
\newcommand{\Zeta}{\Zeta}
\newcommand{\kappa}{\kappa}
\newcommand{\Kappa}{\Kappa}
\renewcommand{\l}{\lambda}
\renewcommand{\L}{\Lambda}
\newcommand{\mu}{\mu}
\newcommand{\Mu}{\Mu}
\newcommand{\nu}{\nu}
\newcommand{\Nu}{\Nu}
\newcommand{\phi}{\phi}
\newcommand{\varphi}{\varphi}
\newcommand{\Phi}{\Phi}
\newcommand{\chi}{\chi}
\newcommand{\Chi}{\Chi}
\newcommand{\pi}{\pi}
\renewcommand{\P}{\Pi}
\newcommand{\varpi}{\varpi}
\newcommand{\rho}{\rho}
\newcommand{\Rho}{\Rho}
\newcommand{\sigma}{\sigma}
\renewcommand{\S}{\Sigma}
\renewcommand{\t}{\tau}
\newcommand{\Tau}{\Tau}
\newcommand{\upsilon}{\upsilon}
\newcommand{\upsilon}{\upsilon}
\renewcommand{\o}{\omega}
\renewcommand{\O}{\Omega}
\newcommand{\bar{X}}{\bar{X}}
\newcommand{\bar{D}}{\bar{D}}
\newcommand{\bar{W}}{\bar{W}}
\newcommand{\bar{\Psi}}{\bar{\Psi}}
\newcommand{\bar{F}}{\bar{F}}
\newcommand{\bar{H}}{\bar{H}}
\newcommand{\bar{P}}{\bar{P}}
\newcommand{\bar{\phi}}{\bar{\phi}}
\newcommand{\bar{\chi}}{\bar{\chi}}
\newcommand{\tilde{{\cal X}}}{\tilde{{\cal X}}}
\newcommand{\tilde{\chi}}{\tilde{\chi}}
\newcommand{\tilde{\eta}}{\tilde{\eta}}
\newcommand{\tilde{G}}{\tilde{G}}
\newcommand{\tilde{{\cal Q}}}{\tilde{{\cal Q}}}
\newcommand{{\cal P}}{{\cal P}}
\newcommand{{\cal Q}}{{\cal Q}}
\newcommand{{\cal X}}{{\cal X}}
\newcommand{\bar{{\cal P}}}{\bar{{\cal P}}}
\newcommand{\tilde{Q}}{\tilde{Q}}
\newcommand{\tilde{\l}}{\tilde{\l}}
\newcommand{\tilde{I}}{\tilde{I}}
\newcommand{\bar{\Phi}}{\bar{\Phi}}
\def\Mb{\kern 2pt\mathchoice
\vbox{\hrule width10pt height 0.4pt depth 0pt
\kern 1.2pt\hbox{\kern -2pt$\displaystyle M$}}}
\vbox{\hrule width10pt height 0.4pt depth 0pt
\kern 1.2pt\hbox{\kern -2pt$\textstyle M$}}}
\vbox{\hrule width6pt height 0.4pt depth 0pt
\kern 1.0pt\hbox{\kern -2pt$\scriptstyle M$}}}
\vbox{\hrule width5pt height 0.4pt depth 0pt
\kern 0.8pt\hbox{\kern -2pt$\scriptscriptstyle M$}}}}
\def\Sb{\kern 2pt\mathchoice
\vbox{\hrule width6pt height 0.4pt depth 0pt
\kern 1.2pt\hbox{\kern -2pt$\displaystyle S$}}}
\vbox{\hrule width6pt height 0.4pt depth 0pt
\kern 1.2pt\hbox{\kern -2pt$\textstyle S$}}}
\vbox{\hrule width3.5pt height 0.4pt depth 0pt
\kern 1.0pt\hbox{\kern -2pt$\scriptstyle S$}}}
\vbox{\hrule width3pt height 0.4pt depth 0pt
\kern 0.8pt\hbox{\kern -2pt$\scriptscriptstyle S$}}}}
\def\Rb{\kern 2pt\mathchoice
\vbox{\hrule width5.5pt height 0.4pt depth 0pt
\kern 1.2pt\hbox{\kern -2.5pt$\displaystyle R$}}}
\vbox{\hrule width5.5pt height 0.4pt depth 0pt
\kern 1.2pt\hbox{\kern -2.5pt$\textstyle R$}}}
\vbox{\hrule width3.5pt height 0.4pt depth 0pt
\kern 1.0pt\hbox{\kern -2.2pt$\scriptstyle R$}}}
\vbox{\hrule width3pt height 0.4pt depth 0pt
\kern 0.8pt\hbox{\kern -2.2pt$\scriptscriptstyle R$}}}}
\def\pp{{\mathchoice
%
\kern 1pt%
\raise 1pt
\vbox{\hrule width5pt height0.4pt depth0pt
\kern -2pt
\hbox{\kern 2.3pt
\vrule width0.4pt height6pt depth0pt
}
\kern -2pt
\hrule width5pt height0.4pt depth0pt}%
\kern 1pt
}
\kern 1pt%
\raise 1pt
\vbox{\hrule width4.3pt height0.4pt depth0pt
\kern -1.8pt
\hbox{\kern 1.95pt
\vrule width0.4pt height5.4pt depth0pt
}
\kern -1.8pt
\hrule width4.3pt height0.4pt depth0pt}%
\kern 1pt
}
\kern 0.5pt%
\raise 1pt
\vbox{\hrule width4.0pt height0.3pt depth0pt
\kern -1.9pt
\hbox{\kern 1.85pt
\vrule width0.3pt height5.7pt depth0pt
}
\kern -1.9pt
\hrule width4.0pt height0.3pt depth0pt}%
\kern 0.5pt
}
\kern 0.5pt%
\raise 1pt
\vbox{\hrule width3.6pt height0.3pt depth0pt
\kern -1.5pt
\hbox{\kern 1.65pt
\vrule width0.3pt height4.5pt depth0pt
}
\kern -1.5pt
\hrule width3.6pt height0.3pt depth0pt}%
\kern 0.5p
}
}}
\def\mm{{\mathchoice
%
%
\kern 1pt
\raise 1pt \vbox{\hrule width5pt height0.4pt depth0pt
\kern 2pt
\hrule width5pt height0.4pt depth0pt}
\kern 1pt}
\kern 1pt
\raise 1pt \vbox{\hrule width4.3pt height0.4pt depth0pt
\kern 1.8pt
\hrule width4.3pt height0.4pt depth0pt}
\kern 1pt}
\kern 0.5pt
\raise 1pt
\vbox{\hrule width4.0pt height0.3pt depth0pt
\kern 1.9pt
\hrule width4.0pt height0.3pt depth0pt}
\kern 1pt}
\kern 0.5pt
\raise 1pt \vbox{\hrule width3.6pt height0.3pt depth0pt
\kern 1.5pt
\hrule width3.6pt height0.3pt depth0pt}
\kern 0.5pt}
}}
\def\pd{{\kern0.5pt
+ \kern-5.05pt \raise5.8pt\hbox{$\textstyle.$}\kern
0.5pt}}
\def\pmd{{\kern0.5pt
\pm \kern-5.05pt \raise6.3pt\hbox{$\textstyle.$}\kern1.5pt}}
\def\md{{\mathchoice
{{\kern 1pt - \kern-6.2pt \raise5pt\hbox{$\textstyle.$}\kern 1pt}}}
{{\kern 1pt - \kern-6.2pt \raise5pt\hbox{$\textstyle.$}\kern 1pt}}}
{\kern0.5pt - \kern-5.05pt \raise3.4pt\hbox{$\textstyle.$}\kern0.5pt}}
{\kern0.5pt - \kern-5.05pt \raise3.4pt\hbox{$\textstyle.$}\kern0.5pt}}}}
\def\rule[-15pt]{0pt}{0pt}{\rule[-15pt]{0pt}{0pt}}
\newcommand{\hat{E}}{\hat{E}}
\newcommand{\check{E}}{\check{E}}
\newcommand{A_+^{~-}}{A_+^{~-}}
\newcommand{A_-^{~+}}{A_-^{~+}}
\newcommand{A_{\pd}^{~\md}}{A_{\pd}^{~\md}}
\newcommand{A_{\md}^{~\pd}}{A_{\md}^{~\pd}}
\newcommand{A_{\pd}^{~\md}}{A_{\pd}^{~\md}}
\newcommand{A_{\md}^{~\pd}}{A_{\md}^{~\pd}}
\newcommand{\hat{C}}{\hat{C}}
\newcommand{\check{C}}{\check{C}}
\newcommand{\Ch_{-\pd}^{~~~\pp}}{\hat{C}_{-\pd}^{~~~\pp}}
\newcommand{\Ch_{+\md}^{~~~\pp}}{\hat{C}_{+\md}^{~~~\pp}}
\newcommand{\Ch_{+\md}^{~~~\mm}}{\hat{C}_{+\md}^{~~~\mm}}
\newcommand{\Ch_{-\pd}^{~~~\mm}}{\hat{C}_{-\pd}^{~~~\mm}}
\newcommand{ \buildrel \leftarrow \over \Ec}{ \buildrel \leftarrow \over \check{E}}
\newcommand{\buildrel \leftarrow \over H}{\buildrel \leftarrow \over H}
\newcommand{ \buildrel \leftarrow \over \Eh}{ \buildrel \leftarrow \over \hat{E}}
\newcommand{ \buildrel \leftarrow \over D}{ \buildrel \leftarrow \over D}
\newcommand{\buildrel \leftrightarrow \over \pa_{\pp}}{\buildrel \leftrightarrow \over \partial_{\pp}}
\newcommand{\buildrel \leftarrow \over \x}{\buildrel \leftarrow \over \chi}
\newcommand{\buildrel \leftrightarrow \over {\cal D}_{\pp}}{\buildrel \leftrightarrow \over {\cal D}_{\pp}}
\newcommand{\buildrel \leftrightarrow \over {\cal D}_{\mm}}{\buildrel \leftrightarrow \over {\cal D}_{\mm}}
\newcommand{{\dot{\alpha}}}{{\dot{\alpha}}}
\newcommand{{\dot{\beta}}}{{\dot{\beta}}}
\newcommand{\nabla}{\nabla}
\newcommand{\bar{\nabla}}{\bar{\nabla}}
\newcommand{\nabla_{+}}{\nabla_{+}}
\newcommand{\nabla_{-}}{\nabla_{-}}
\newcommand{\nabla_{\pd}}{\nabla_{\pd}}
\newcommand{\nabla_{\md}}{\nabla_{\md}}
\newcommand{\nabla_{\pp}}{\nabla_{\pp}}
\newcommand{\nabla_{\mm}}{\nabla_{\mm}}
\newcommand{\nabla^2}{\nabla^2}
\newcommand{{\bar{\nabla}}^2}{{\bar{\nabla}}^2}
\newcommand{\nabla_{\alpha}}{\nabla_{\alpha}}
\newcommand{\nabla_{\dot{\alpha}}}{\nabla_{\dot{\alpha}}}
\newcommand{\nabla_A}{\nabla_A}
\newcommand{{\tilde{\Del}}^2}{{\tilde{\nabla}}^2}
\newcommand{D_{\pd}}{D_{\pd}}
\newcommand{D_{\md}}{D_{\md}}
\newcommand{D_{\pp}}{D_{\pp}}
\newcommand{D_{\mm}}{D_{\mm}}
\newcommand{{\bf D}}{{\bf D}}
\newcommand{{\bf D}_+}{{\bf D}_+}
\newcommand{{\bf D}_-}{{\bf D}_-}
\newcommand{{\bf D}_{\pd}}{{\bf D}_{\pd}}
\newcommand{{\bf D}_{\md}}{{\bf D}_{\md}}
\newcommand{{\bf D}_{\pp}}{{\bf D}_{\pp}}
\newcommand{{\bf D}_{\mm}}{{\bf D}_{\mm}}
\newcommand{{\bf D}_{\alpha}}{{\bf D}_{\alpha}}
\newcommand{{\bf D}_{\dot{\alpha}}}{{\bf D}_{\dot{\alpha}}}
\newcommand{{\bf D}_A}{{\bf D}_A}
\newcommand{{\cal D}}{{\cal D}}
\newcommand{{\cal D}_+}{{\cal D}_+}
\newcommand{{\cal D}_-}{{\cal D}_-}
\newcommand{{\cal D}_\pd}{{\cal D}_\pd}
\newcommand{{\cal D}_\md}{{\cal D}_\md}
\newcommand{{\cal D}_\pp}{{\cal D}_\pp}
\newcommand{{\cal D}_\mm}{{\cal D}_\mm}
\newcommand{{\cal D}_\alpha}{{\cal D}_\alpha}
\newcommand{{\cal D}_{\dot{\alpha}}}{{\cal D}_{\dot{\alpha}}}
\newcommand{{\cal D}_A}{{\cal D}_A}
\def\scriptstyle{\scriptstyle}
\def\scriptscriptstyle{\scriptscriptstyle}
\newcommand{\reff}[1]{(\ref{#1})}
\parskip 0.3cm
\newcommand{{\frac{1}{2}}}{{\frac{1}{2}}}
\newcommand{{\frac{i}{2}}}{{\frac{i}{2}}}
\newcommand{\frac{1}{2}}{\frac{1}{2}}
\newcommand{\frac{i}{2}}{\frac{i}{2}}
\newcommand{{\scriptstyle\frac{1}{4}}}{{\scriptstyle\frac{1}{4}}}
\newcommand{\bl}[1]{\Bigl(#1\Bigr)}
\newcommand{\blt}[1]{\Bigl( #1\Bigr)_H}
\newcommand{\cbl}[1]{\Bigl\{ #1\Bigr\}_H}
\defH{H}
\def\widetilde{m}{\widetilde{m}}
\def\mbox{sh}{\mbox{sh}}
\def\mbox{ch}{\mbox{ch}}
\newcommand{\NP}[1]{Nucl.\ Phys.\ {\bf #1}}
\newcommand{\PL}[1]{Phys.\ Lett.\ {\bf #1}}
\newcommand{\NC}[1]{Nuovo Cimento {\bf #1}}
\newcommand{\CMP}[1]{Comm.\ Math.\ Phys.\ {\bf #1}}
\newcommand{\PR}[1]{Phys.\ Rev.\ {\bf #1}}
\newcommand{\PRL}[1]{Phys.\ Rev.\ Lett.\ {\bf #1}}
\newcommand{\MPL}[1]{Mod.\ Phys.\ Lett.\ {\bf #1}}
\renewcommand{\thefootnote}{\fnsymbol{footnote}}
\begin{document}
\newpage
\begin{titlepage}
\begin{flushright}
{hep-th/9508139}\\
{BRX-TH-378}
\end{flushright}
\vspace{2cm}
\begin{center}
{\bf {\large SUPERSPACE MEASURES, INVARIANT ACTIONS, AND COMPONENT
PROJECTION FORMULAE FOR (2,2)
SUPERGRAVITY}}\\
\vspace{1.5cm}
Marcus T. Grisaru\footnote{
Work partially supported by the National Science Foundation under
grant PHY-92-22318.} \\
and\\
\vspace{1mm}
Marcia E. Wehlau\footnote{\hbox to \hsize{Current address:
Mars Scientific Consulting, 28 Limeridge Dr.,
Kingston, ON CANADA K7K~6M3}}\\
\vspace{1mm}
{\em Physics Department, Brandeis University, Waltham, MA 02254, USA}
\vspace{1.1cm}
{{ABSTRACT}}
\end{center}
\begin{quote}
In the framework of the prepotential description of superspace
two-dimensional $(2,2)$ supergravity,
we discuss the construction of invariant integrals. In addition to the full
superspace
measure, we derive the measure for chiral superspace, and obtain the
explicit
expressions for going from superspace actions to component actions.
We consider both the minimal $U_A(1)$ and the extended
$U_V(1) \otimes U_A(1)$
theories.
\end{quote}
\vfill
\begin{flushleft}
August 1995
\end{flushleft}
\end{titlepage}
\newpage
\renewcommand{\thefootnote}{\arabic{footnote}}
\setcounter{footnote}{0}
\newpage
\pagenumbering{arabic}
\section{Introduction}
Recently \cite{MGMW}, we have described the solution of the
constraints
for $(2,2)$ supergravity
in $(2,2)$ superspace in terms of unconstrained prepotentials.
For the
nonminimal $U_V(1)
\otimes U_A(1)$ theory, these are a real vector superfield $H^m$
and a
general
scalar superfield $S$. In the ``degauged'' minimal theory where one
of the
$U(1)$ connections
is set to zero, a further constraint expresses the superfield $S$ in terms
of a
chiral superfield $\sigma$ (for the $U_A(1)$ theory), or a twisted chiral
superfield
$\tilde{\sigma}$
(for the $U_V(1)$ theory). We gave explicit expressions for the
vielbein, the
connections, and the vielbein determinant $E$. With these results
at hand,
several
applications become possible. In particular, we have studied the
theory in
light-cone gauge and derived the Ward identities associated with the
(nonlocal)
induced $(2,2)$ supergravity action \cite{lc}. One can also
discuss the
fully supersymmetric quantization of the theory. We have described
in a
separate work the general coupling of two-dimensional $(2,2)$
supersymmetric matter to supergravity and the corresponding
component
actions \cite{JGMGMW}.
For this latter application, a knowledge of invariant superspace
measures
in full
superspace and in chiral (or twisted chiral) superspace is
necessary, as
well as projection
formulae that allow one to obtain the corresponding component
actions.
In this paper we describe this aspect of the theory -- the
construction of
invariant actions, the determination of the measures (in particular
of chiral
or twisted
chiral densities), and the projection formulae that enable us to go from
superspace actions
to component actions.
In the words of an esteemed colleague ``the construction of
superinvariants
is more an art
than a science". We show here that, at least in the two-dimensional
$(2,2)$ theory,
this construction is in fact a science.
Our paper is organized as follows: in section 2 we summarize the
relevant information
about the $(2,2)$ supergravity theory and the solution of the
constraints. In section 3
we show how, for the minimal $U_A(1)$ theory, one obtains the
chiral measure for
integration over chiral superspace (or, equivalently, the twisted
chiral results for the
$U_V(1)$ theory). In section 4, we derive the component projection
formula for
going from full superspace (or chiral superspace) to components.
In the same section
we
obtain an alternative projection formula which involves at an
intermediate
stage a twisted chiral projector rather than the chiral projector.
In section 5, we show
how to express the covariant derivatives of the $U_V(1) \otimes
U_A(1)$ theory in terms
of those of the degauged $U_A(1)$ theory, and derive the chiral
measure for the
former. Section 6 contains
our conclusions. We use the conventions of refs. \cite{MGMW,lc}.
Appendix A
contains the definition of the supergravity component fields in
Wess-Zumino gauge,
and additional information about these derivatives. Appendices B
and C contain some
details of the derivations in the main text.
\section{The (2,2) constraints and their solution}
We summarize in this section the main results of refs. \cite{MGMW,lc}.
Tangent
space Lorentz, $U_V(1)$ and $U_A(1)$ generators,
denoted here by ${\cal M}$, ${\cal Y}$ and
${\cal Y}'$ respectively, are defined by their action on spinors:
\begin{eqnarray}
[~{\cal M} \, , \, \psi_{\pm}~] ~=~ \pm \frac{1}{2}
\psi_{\pm} ~~~~&,&~~~~
[~{\cal M} \, , \, \psi_{\pmd}~] ~=~ \pm \frac{1}{2}
\psi_{\pmd} ~~~,
\nonumber\\
{[}~{\cal Y} \, , \, \psi_{\pm}~] ~=~ - \frac{i}{2}
\psi_{\pm} ~~~~&,&~~~~
[~ {\cal Y} \, , \, \psi_{\pmd}~] ~=~ + \frac{i}{2}
\psi_{\pmd} ~~~,
\nonumber\\
{[}~ {\cal Y}' \, , \, \psi_{\pm}~] ~=~ \mp \frac{i}{2}
\psi_{\pm}~~~~&,&~~~~
[~ {\cal Y}' \, , \, \psi_{\pmd}~ ] ~=~ \pm \frac{i}{2}
\psi_{\pmd} ~~~ .
\end{eqnarray}
It is also useful to define the combinations
\begin{eqnarray}
M &=& \frac{1}{2}({\cal M} +i {\cal Y} ')~~~~,~~~
\bar{M}= \frac{1}{2}({\cal M} -i {\cal Y} ') \non\\
N &=& \frac{1}{2}({\cal M} +i {\cal Y} )~~~~,~~~
\bar{N}= \frac{1}{2}({\cal M} -i {\cal Y} )~~.
\end{eqnarray}
The covariant derivatives are defined by
\begin{eqnarray}
{\nabla}_A &=& E_A + \Phi_A {\cal M} + \S_A' {\cal Y}'+ \S_A
{\cal Y} \non\\
&=& E_A + \O_A M + \Gamma_A \Mb + \S_A {\cal Y} ~~.
\end{eqnarray}
They satisfy the constraints which define the 2D, N = 2 $U_V (1)
\otimes U_A (1)$
supergravity,
\begin{eqnarray}
\{ { \nabla}_+ ~,~ { \nabla}_+ \} ~&=&~ 0 ~~~~~~~~~~~,~~~~~~
\{ { \nabla}_-
{}~,~ { \nabla}_- \} ~=~ 0 \nonumber\\
\{ { \nabla}_+ ~,~ { \nabla}_{\pd} \} ~&=&~ i { \nabla}_{\pp}
~~~~~~~,~~~~~~~ \{ { \nabla}_- ~,~ { \nabla}_{\md} \} ~=~ i
{ \nabla}_{\mm} ~~~ \nonumber\\
\{ { \nabla}_+ ~,~ { \nabla}_- \} ~&=&~ - \, {\bar R} \,
{\bar M} ~~~~~,~~~~~ \{ { \nabla}_+ ~,~ { \nabla}_{\md} \} ~=~ - \,
{\bar F} \, {\bar N} ~~~.
\end{eqnarray}
{}From the constraints follow the additional commutators
\begin{eqnarray}
{[}~{ {\nabla}}_{+} \, , \, { {\nabla}}_{\pp} ~] &=& 0 ~~~, ~~~[~
{ {\nabla}}_{-} \, , \, { {\nabla}}_{\mm} ~] = 0
{}~~~, \\ \non
{[}~{ {\nabla}}_{\pd} \, , \, { {\nabla}}_{\pp} ~] &=& 0 ~~~, ~~~[~
{ {\nabla}}_{\md} \, , \, { {\nabla}}_{\mm} ~] =
0 ~~~, \\ \non
{[}~{ {\nabla}}_{+} \, , \, { {\nabla}}_{\mm} ~] &=& - {\frac{i}{2}} \Rb
{ {\nabla}}_{\md} -i({ {\nabla}}_{\md} \Rb) \Mb
- {\frac{i}{2}} \bar{F} {\nabla}_-
-i ({ {\nabla}}_- \bar{F} ) \bar{N}
{}~~~, \\ \non
{[}~{ {\nabla}}_{\pd} \, , \, { {\nabla}}_{\mm} ~] &=& {\frac{i}{2}} R
{ {\nabla}}_{-} +i( { {\nabla}}_- R) M
+ {\frac{i}{2}} F
{ {\nabla}}_{\md} +i( { {\nabla}}_{\md} F) N ~~~, \\ \non
{[}~{ {\nabla}}_{-} \, , \, { {\nabla}}_{\pp} ~] &=& {\frac{i}{2}} \Rb
{ {\nabla}}_{\pd} -i({ {\nabla}}_{\pd} \Rb)\Mb
+{\frac{i}{2}} F
{ {\nabla}}_{+} -i({ {\nabla}}_{+} F)N
{}~~~, \\ \non
{[}~{ {\nabla}}_{\md} \, ,\, { {\nabla}}_{\pp} ~] &=& - {\frac{i}{2}} R
{ {\nabla}}_{+} +i({ {\nabla}}_{+} R) M
-{\frac{i}{2}} \bar{F}
{ {\nabla}}_{\pd} +i({ {\nabla}}_{\pd} \bar{F}) \bar{N}
~~~,
\end{eqnarray}
and
\newpage
\begin{eqnarray}
{[}~ { {\nabla}}_{\pp} \, , \, { {\nabla}}_{\mm} ~] &=& \frac{1}{2} (
{ {\nabla}}_+ R) { {\nabla}}_- +\frac{1}{2} ({ {\nabla}}_- R) { {
\nabla}}_+ - \frac{1}{2} ({ {\nabla}}_{\pd} \Rb) { {\nabla}}_{\md} - \frac{1}{2} (
{ {\nabla}}_{\md} \Rb) { {\nabla}}_{\pd} \non \\
&& - \frac{1}{2} R \Rb \Mb - \frac{1}{2} R \Rb M + ({{ {\nabla}}}^2 R) M -(
{{ {\bar{\nabla}}}}^2 \Rb) \Mb \non \\
&& + \frac{1}{2} (
{ {\nabla}}_+ F) { {\nabla}}_{\md} +\frac{1}{2} ({ {\nabla}}_{\md} F)
{ {
\nabla}}_+ - \frac{1}{2} ({ {\nabla}}_{\pd} \bar{F}) { {\nabla}}_{-}
- \frac{1}{2} (
{ {\nabla}}_{-} \bar{F}) { {\nabla}}_{\pd} \non \\
&& - \frac{1}{2} F \bar{F} \bar{N} - \frac{1}{2} F \bar{F} N +
({ {\nabla}}_+ { {\nabla}}_{\md} F) N -(
{ {\nabla}}_{\pd} { {\nabla}}_- \bar{F}) \bar{N}
\label{DelppDelmm} ~~.
\end{eqnarray}
Furthermore,
for the minimal supergravities one restricts the gauge group
so that either
$F=0$ for the $U_A(1)$ version,
or $R=0$ for the $U_V(1)$ version, by ``degauging'', i.e. by
setting either
$\Sigma_{A}=0$ or
${\S}_{A}'=0$.
The solution of the constraints is obtained in terms of the ``hat''
differential operators
\begin{equation}
\hat{E}_{\pm}= e^{-H}D_{\pm}e^{H}~~~~,~~~~ \hat{E}_{\pmd}
= e^{H} D_{\pmd}
e^{-H}
\end{equation}
with $H=H^mi\partial_m$, where $H^m$ is a real vector superfield.
The spinorial vielbein is expressed in terms of these operators,
as well as an
additional (superscale and $U(1)$ compensator) general complex
superfield $S$
as
\begin{eqnarray}
E_+ \equiv e^{\Sb}(\hat{E} _+ + A_+^{~-} \hat{E}_-) ~~~~&,&~~~~E_- \equiv
e^{\Sb}
(\hat{E}_-+A_-^{~+} \hat{E}_+) \non\\
E_{\pd} \equiv e^S(\hat{E} _{\pd} + A_{\pd}^{~\md} \hat{E}_\md)
~~~~&,&~~~~E_\md
\equiv e^S
(\hat{E}_\md+A_{\md}^{~\pd}
\hat{E}_\pd) ~~.
\end{eqnarray}
The vectorial vielbein is calculated from the constraints.
The $A$'s, as well as the vielbein determinant $E$ are given explicitly
in ref. \cite{MGMW}.
We emphasize that they are functions of the superfield $H^m$
{\em only}.
The connections were derived in ref. \cite{MGMW}, and we list them
here for convenience:
\begin{eqnarray}
{\Omega}_+&=&+ e^{{\Sb}}(\hat{E}_-{A_+^{~-}} - {A_+^{~-}}\hat{E}_+ {A_-^{~+}} )\non\\
{\Omega}_- &=&-e^{{\Sb}}( \hat{E}_+{A_-^{~+}} -{A_-^{~+}} \hat{E}_-{A_+^{~-}}) \non\\
{\Sigma}_+&=&-2ie^{{\Sb}}(\hat{E}_+{\Sb }+{A_+^{~-}}\hat{E}_-{\Sb} )-ie^{{\Sb}}
(\hat{E}_-{A_+^{~-} } -{A_+^{~-}}\hat{E}_+{A_-^{~+}} ) \label{connections} \\
{\Sigma}_-&=&-2ie^{{\Sb}}(\hat{E}_-{\Sb} +{A_-^{~+}} \hat{E}_+{\Sb} )-ie^{{\Sb}}
(\hat{E}_+{A_-^{~+}} -{A_-^{~+}}\hat{E}_-{A_+^{~-}} ) \nonumber\\
{\Gamma}_+ &=& +2e^{{\Sb}}(\hat{E}_+S+{A_+^{~-} }\hat{E}_-{S})+2e^{{\Sb}}
(\hat{E}_+{\Sb}
+{A_+^{~-}} \hat{E}_-{\Sb})+e^{{\Sb}}(\hat{E}_-{A_+^{~-}} -{A_+^{~-}} \hat{E}_+{A_-^{~+}} ) \non\\
{\Gamma}_- &=& -2e^{{\Sb}}(\hat{E}_-{S}+{A_-^{~+}} \hat{E}_+{S}-2e^{{\Sb}}
(\hat{E}_-{\Sb}
+{A_-^{~+}} \hat{E}_+{\Sb})-e^{{\Sb}}(\hat{E}_+{A_-^{~+}} -{A_-^{~+}} \hat{E}_-{A_+^{~-} }) \non
\end{eqnarray}
The $U_V(1) \otimes U_A(1)$ theory contains in addition to the
irreducible
supergravity multiplet a vector multiplet. A minimal theory is obtained by
setting one
of the field strengths $R$ or $F$ (or the corresponding connections)
to zero.
In ref. \cite{MGMW} we
worked out the implications of the additional constraint $F=0$ for
the minimal
$U_A(1)$
theory and found that the superfield $S$ satisfies a constraint that
expresses
it
in terms of an arbitrary covariantly chiral superfield $\sigma$ as follows:
\begin{equation}
e^S = e^{{\sigma}} \frac{ \left[1\cdot e^{\buildrel \leftarrow \over H}
\right]^{-\frac{1}{2}}}{[1-A_{\pd}^{~\md} A_{\md}^{~\pd}]^{\frac{1}{2}}}
E^{-\frac{1}{2}}~~~~,
{}~~~~e^{\Sb} = e^{\bar{\sigma}} \frac{ \left[1\cdot e^{-\buildrel \leftarrow \over H}
\right]^{-\frac{1}{2}}}{[1-A_+^{~-} A_-^{~+}]^{\frac{1}{2}}} E^{-\frac{1}{2}}~~,
\label{degauge}
\end{equation}
where $\buildrel \leftarrow \over H$ indicates
that the differential operator in $H^m i \partial_m$ acts on objects
to its left.
The unconstrained real vector superfield $H^m$ and the chiral
scalar
superfield $\sigma$ are the prepotentials of the minimal
$U_A(1)$ $(2,2)$
supergravity. The mirror image $U_V(1)$ theory is obtained
by interchanging
$-$ and $\md ~$ indices everywhere (and interchanging $R$
with $F$, and
$M$ with $N$), and amounts
to replacing the chiral superfield $\sigma$ by a twisted chiral superfield
$\tilde{\sigma}$.
In the following, we shall concentrate primarily on the minimal
$U_A(1)$ theory.
\sect{From full superspace measure to chiral superspace measure}
In full superspace, invariant actions are constructed by means of the
vielbein
determinant as
\begin{equation}
{\cal S} = \int d^2x d^4 \theta E^{-1} {\cal L}
\end{equation}
where ${\cal L}$ is an arbitrary scalar function of superfields. The
invariance
of
this integral under superspace coordinate transformations can be
established by
standard means; see for example {\em Superspace}, sect. 5.5
cite{Superspace}.
In addition \cite{Superspace,lc}, one can show that for superspace
transformations under which
covariantly chiral and antichiral superfields transform as
\begin{equation}
\Phi \rightarrow e^{i\L}\Phi ~~~~,~~~ \bar{\Phi} \rightarrow
e^{i \bar{\L}}
\bar{\Phi}
\end{equation}
with
\begin{eqnarray}
\L&=&\L^m i\partial_m +\L^{\a}iD_{\a} +\L^{{\dot{\alpha}}}iD_{{\dot{\alpha}}} \non\\
\bar{\L}&=&\bar{\L}^m i\partial_m +\bar{\L}^{\a}iD_{\a} +
\bar{\L}^{{\dot{\alpha}}}iD_{{\dot{\alpha}}} ~~,
\end{eqnarray}
the prepotential $H^a$ must transform as
\begin{equation}
e^{2H} \rightarrow e^{i\bar{\L}} e^{2H} e^{-i\L}
\end{equation}
and the vielbein determinant transforms as
\begin{equation}
E^{-1} e^{\buildrel \leftarrow \over H} \rightarrow E^{-1}e^{\buildrel \leftarrow \over H} e^{i \buildrel
\leftarrow \over
{\bar{\L}}} ~~.
\end{equation}
In the minimal $U_A(1)$ theory we show now that one can
rewrite the full superspace integral as an integral over chiral
superspace, with an appropriate measure ${\cal E}$ that we
determine
explicitly:
\begin{equation}
{\cal S}= \int d^2x d^2 \theta {\cal E}^{-1} \bar{\nabla}^2 {\cal L}|_{\bar{\theta}=0}
\end{equation}
with
\begin{equation}
{\cal E}^{-1} = e^{-2\sigma} (1.e^{\buildrel \leftarrow \over H}) ~~.
\end{equation}
We proceed by deriving first an explicit expression for $\bar{\nabla}^2$
acting
on an {\em arbitrary} scalar $L$.
We start with
\begin{eqnarray}
{\bar{\nabla}}^2 L&=& \nabla_{\pd} \nabla_{\md} L \non \\
&=& [e^S(\hat{E}_\pd + A_{\pd}^{~\md} \hat{E}_\md) - \frac{1}{2} \O_\pd]
[e^S(\hat{E}_\md + A_{\md}^{~\pd} \hat{E}_\pd)] L ~~.
\end{eqnarray}
{} From the complex conjugate of equations \reff{connections},
with $\S_\pd
= 0$,
we have
\begin{equation}
\O_\pd = -2e^S(\hat{E}_\pd S + A_{\pd}^{~\md} \hat{E}_\md S) ~~.
\end{equation}
Using this we rewrite ${\bar{\nabla}}^2 L$ as
\begin{eqnarray}
{\bar{\nabla}}^2 L &=& (\hat{E}_\pd + A_{\pd}^{~\md} \hat{E}_\md)
e^{2S}(\hat{E}_\md + A_{\md}^{~\pd} \hat{E}_\pd) L ~\nonumber\\
&=&{(\hat{E}_\pd + A_{\pd}^{~\md} \hat{E}_\md) (\hat{E}_\md + A_{\md}^{~\pd}
\hat{E}_\pd) e^{2S} L} \non \\
&&-2 (\hat{E}_\pd + A_{\pd}^{~\md} \hat{E}_\md)(\hat{E}_\md S + A_{\md}^{~\pd} \hat{E}_\pd S)
e^{2S} L ~~.
\end{eqnarray}
Again from the complex conjugate of equations \reff{connections},
setting
$\S_\md = 0$, we obtain
\begin{equation}
2(\hat{E}_\md S + A_{\md}^{~\pd} \hat{E}_\pd S) = -(\hat{E}_\pd A_{\md}^{~\pd} - A_{\md}^{~\pd}
\hat{E}_\md A_{\pd}^{~\md}) ~~,
\end{equation}
which, when substituted into the expression for ${\bar{\nabla}}^2 L$ above,
gives
\begin{eqnarray}
{\bar{\nabla}}^2 L&=& [(1-A_{\pd}^{~\md} A_{\md}^{~\pd})\hat{E}_\pd \hat{E}_\md + \hat{E}_\md(A_{\pd}^{~\md}
A_{\md}^{~\pd})\hat{E}_\pd - \hat{E}_\pd(A_{\pd}^{~\md} A_{\md}^{~\pd})\hat{E}_\md \non \\
&& - \hat{E}_\pd \hat{E}_\md (A_{\pd}^{~\md} A_{\md}^{~\pd})] e^{2S} L \non\\
&=& \hat{E}_\pd \hat{E}_\md [(1-A_{\pd}^{~\md} A_{\md}^{~\pd})e^{2S} L]~~.
\end{eqnarray}
We have used the complex conjugates of (3.5) and (3.6) of ref.
\cite{MGMW}, namely,
\begin{eqnarray}
\hat{E}_\pd A_{\pd}^{~\md} + A_{\pd}^{~\md} \hat{E}_\md A_{\pd}^{~\md} &=& 0 \non \\
\hat{E}_\md A_{\md}^{~\pd} + A_{\md}^{~\pd} \hat{E}_\pd A_{\md}^{~\pd} &=& 0
\end{eqnarray}
and
\begin{eqnarray}
\hat{E}_\pd \hat{E}_\md A_{\pd}^{~\md} &=& 0 \non \\
\hat{E}_\md \hat{E}_\pd A_{\md}^{~\pd} &=& 0 ~~.
\end{eqnarray}
We obtain, using \reff{degauge}
\begin{equation}
{\bar{\nabla}}^2 L = \hat{E}_\pd \hat{E}_\md [e^{2 \sigma} (1.e^{\buildrel \leftarrow \over H})^{-1}E^{-1} L] ~~.
\label{ident}
\end{equation}
We define $ L'$ by $L = e^{-2 \sigma}(1.e^{\buildrel \leftarrow \over H}) L'$, so that
\reff{ident} becomes
\begin{eqnarray}
{\bar{\nabla}}^2 [e^{-2 \sigma}(1.e^{\buildrel \leftarrow \over H}) L']
&=& \hat{E}_\pd \hat{E}_\md (E^{-1} L') \non \\
&=& e^H \bar{D}^2 e^{-H}(E^{-1} L') ~~.
\end{eqnarray}
Multiplying both sides by $e^{-H}$ and integrating over chiral
superspace, we
obtain
\begin{equation}
\int d^2x d^2 \theta \bar{D}^2 e^{-H} E^{-1} L' =\int d^2x d^2 \theta e^{-H}
e^{-2 \sigma} (1.e^{\buildrel \leftarrow \over H})L' ~~.
\end{equation}
Integrating the operator $e^{-H}$ by parts on both sides
of the equation
gives
\begin{equation}
\int d^2x d^2 \theta \bar{D}^2 [(1.e^{\buildrel \leftarrow \over H})
E^{-1} L'] = \int d^2x d^2 \theta
(1.e^{\buildrel \leftarrow \over H}) {\bar{\nabla}}^2 [e^{-2 \sigma}(1.e^{\buildrel \leftarrow \over H}) L'] ~~.
\end{equation}
Finally, defining ${\cal L}$ by $ L' = (1.e^{\buildrel \leftarrow \over H})^{-1} {\cal L}$, we
obtain
\begin{equation}
\int d^2x d^2 \theta \bar{D}^2 [E^{-1} {\cal L}] =\int d^2x d^2 \theta
e^{-2 \sigma}(1.e^{\buildrel \leftarrow \over H}) {\bar{\nabla}}^2 {\cal L} ~~.
\end{equation}
We have used the fact that $\sigma$ is covariantly chiral to pull the factor
$e^{-2 \sigma}$ past the covariant derivatives.
Since one can go from a full superspace integral to a $d^2 \theta$
integral by
\begin{equation}
\int d^2x d^4 \theta E^{-1} {\cal L} = \int d^2x d^2 \theta \bar{D}^2 [E^{-1}
{\cal L}] |_{\bar{\theta}
=0}
\end{equation}
the relation above leads us to the desired result (and determines the
chiral
measure ${\cal E}$):
\begin{equation}
\int d^2x d^4 \theta E^{-1} {\cal L} = \int d^2 x d^2 \theta e^{-2\sigma}
(1.e^{\buildrel \leftarrow \over H})
\bar{\nabla}^2 {\cal L} |_{\bar{\theta}=0}
~~. \label{chint}
\end{equation}
By construction, the above integral is an invariant. We can replace
$\bar{\nabla}^2 {\cal L} $
by any covariantly chiral expression ${\cal L}_{chiral}$, transforming
under
superspace coordinate
transformations as
\begin{equation}
{\cal L}_{chiral} \rightarrow e^{i {\L}}{\cal L}_{chiral}
\end{equation}
and deduce the transformation properties of the measure
\begin{equation}
e^{-2\sigma} \left( 1\cdot e^{\buildrel \leftarrow \over H} \right)\rightarrow e^{-2\sigma}
\left( 1\cdot e^{\buildrel \leftarrow \over H}
\right)
e^{i \buildrel \leftarrow \over \L} ~~.
\end{equation}
Obviously, a similar result holds for the decomposition to an
antichiral
integral.
In the $U_V(1)$ theory corresponding results, in terms of the
twisted chiral
compensator, can be obtained for twisted chiral integrals of the type
$
\int d^2x d \theta^+ d\theta^{\md}
$.
\sect{From full superspace measure to component integrals}
\subsection{The chiral density projection formula}
We have shown in the previous section that it is possible to
express the full
superspace integral in terms of a chiral integral. In the next step of the
projection
we rewrite this as a component integral over ordinary $d^2x$
space.
By dimensional arguments, and from an examination of the index
structure of
the possible
terms, the density formula for a general $\cal L$ must take the form
\begin{eqnarray}
\int d^2 x d^4 \theta E^{-1} {\cal L} &=& \int d^2x d^2 \theta {\cal E}^{-1}
\bar{\nabla}^2{\cal L}
|_{\bar{\theta}=0}
\non \\
&=& \int d^2 x e^{-1} [\nabla^2 + X^+ \nabla_+ + X^- \nabla_- +Y]
\bar{\nabla}^2
{\cal L} |_{\theta = \bar{\theta}=0}
~~ , \label{gendens}
\end{eqnarray}
where $e$ is the ordinary space zweibein determinant and
$X^+$, $X^-$, and
$Y$ are to be determined. In the past, the determination of the
coefficients $X$, $Y$, has been done essentially by requiring
that the final
component action
be invariant under supersymmetry transformations.
Our derivation is based on the idea that we should obtain the
same result
for the component
action whether we go through the intermediate stage of a chiral
integral,
as in the
equation above, or through a corresponding
antichiral integral.
We illustrate the procedure by using for ${\cal L}$ the
free lagrangian for the chiral multiplet.
As we shall see, proceeding from the above expression, we
obtain a result which is not automatically symmetric in the auxiliary
fields
$F$, $\bar{F}$
of the chiral multiplet. Requiring that the result be symmetric leads to a
unique determination
of the coefficients $X$, $Y$.
We consider the kinetic action for the chiral multiplet
\begin{eqnarray}
\int d^2 x d^4 \theta E^{-1} \bar{\Phi}\Phi
&=& \int d^2 x e^{-1} [\nabla^2 + X ^\a \nabla_\a + Y] {\bar{\nabla}}^2
(\bar{\Phi}\Phi)| \non \\
&=& \int d^2 x e^{-1}[(\nabla^2 {\bar{\nabla}}^2 \bar{\Phi}) \Phi| + (\nabla_{+} {\bar{\nabla}}^2
\bar{\Phi})
(\nabla_{-} \Phi)|
- (\nabla_{-} {\bar{\nabla}}^2 \bar{\Phi})(\nabla_{+} \Phi)| \non \\
&& + ({\bar{\nabla}}^2 \bar{\Phi}) (\nabla^2 \Phi)| + X^+ \nabla_{+} ({\bar{\nabla}}^2 \bar{\Phi} \Phi)
|\non \\
&& + X^- \nabla_{-} ({\bar{\nabla}}^2 \bar{\Phi} \Phi) |+ Y({\bar{\nabla}}^2 \bar{\Phi}) \Phi|]
\label{kinetic}
\end{eqnarray}
according to the density formula.
The (covariant) components of the chiral multiplet are defined by
\begin{eqnarray}
\Phi| = \phi &,& \bar{\Phi}| = \bar{\phi} \nonumber\\
\nabla_{+} \Phi | = \psi _+ &,& \nabla_{\pd} \bar{\Phi}| = \psi_\pd \nonumber \\
\nabla_{-} \Phi | = \psi_- &,& \nabla_{\md} \bar{\Phi}| = \psi_\md \nonumber \\
\nabla^2\Phi| = -iF &,& \bar{\nabla}^2\bar{\Phi}|
=-i \bar{F}
.\end{eqnarray}
supergravity
Appendix A.
In Appendix B we present the complete expressions for the other
component
quantities (involving additional derivatives of $\Phi$)
appearing above.
For our present purposes, we find that the unknowns can be
determined just by
looking at the terms in those expressions that contain the auxiliary
fields, $F$, $\bar{F}$.
We list the relevant terms here:
\begin{eqnarray}
\nabla_{+} {\bar{\nabla}}^2 \bar{\Phi}| &\sim& \psi_\pp^\pd \bar{F}\\
\nabla_{-} {\bar{\nabla}}^2 \bar{\Phi}| &\sim& \psi_\mm^\md \bar{F}\\
\nabla^2 {\bar{\nabla}}^2 \bar{\Phi}| &\sim&
-i( \psi_\pp^\md \psi_\mm^\pd \bar{F} - \psi_\pp^\pd \psi_\mm^\md
\bar{F})
-\frac{i}{2} \bar{B} \bar{F} ~~,
\end{eqnarray}
where we have also used the definition $R| = B $, $ \Rb| = \bar{B}$ .
Substituting into \reff{kinetic}, the sum of the terms
containing auxiliary fields is
\begin{eqnarray}
&&i \psi_\pp^\pd \psi_\mm^\md \bar{F} \phi -i \psi_\pp^\md
\psi_\mm^\pd
\bar{F} \phi
+ \psi_\pp^\pd \bar{F} \psi_- - \psi_\mm^\md \bar{F} \psi_+ \non \\
&-& {\frac{i}{2}} \bar{B} \bar{F} \phi - \bar{F} F + X^+ \psi_\pp^\pd \bar{F} \phi
-i X^+ \bar{F} \psi_+ -i X^- \bar{F} \psi_- \non \\
&+& X^- \psi_\mm^\md \bar{F} \phi -i Y \bar{F} \phi ~~.
\end{eqnarray}
Clearly, except for the $\bar{F} F$ term, this expression is not
symmetric in
barred and unbarred
quantities, and will not agree with the result we would obtain
by going
through
the intermediate stage of an antichiral integral unless we set to
zero the
coefficients of the $\bar{F}\phi
$ and the
$\bar{F}\psi_{\pm}$
terms. We obtain
\begin{equation}
X^+ = i \psi_\mm^\md ~~,~~ X^- = -i \psi_\pp^\pd ~~,~~
Y = - \frac{1}{2} \bar{B} - \psi_\pp^\md \psi_\mm^\pd + \psi_\mm^\md
\psi_\pp^\pd~~,
\end{equation}
leaving $\bar{F} F$ as the only contribution. The form of the chiral
density
formula is
thus determined.
We have, for the final result,
the chiral density projection formula
\begin{eqnarray}
&& \int d^2 x d^4 \theta E^{-1} {\cal L} \non \\
&=& \int d^2 x e^{-1} [ \nabla^2 +i \psi_=^{\dot{-}}\nabla_+ -i
\psi_{\pp}^{\pd}\nabla_-
+(- \frac{1}{2} \bar{B} -\psi_{\pp}^{\dot{-}}\psi_=^{\pd}
+\psi_=^{\dot{-}}\psi_{\pp}^{\pd})]\bar{\nabla}^2
{\cal L} | ~~. \label{chiral density}
\end{eqnarray}
This component expression can be rewritten in terms of a twisted
chiral
projector as discussed in the next subsection.
\subsection{The twisted chiral density projection formula}
In the previous section we have written down the density formula
for going
from a
full superspace integral to a component expression containing as
an intermediate
ingredient a chiral integrand $\bar{\nabla} ^2 {\cal L}$. In this section
we will
derive a
similar formula involving a {\em twisted} chiral integrand
$\nabla_{\pd} \nabla_-
{\cal L}$,
\begin{eqnarray}
&& \int d^2 x d^4 \theta E^{-1} {\cal L} \non \\
&=&\int d^2 x e^{-1}[ \nabla_{\md}\nabla_+ +i \psi^{\pd}_{\pp}
\nabla_{\md}
-i\psi^{-}_{\mm} \nabla_{+} +(\psi^{\pd}_{\pp}\psi^-_{\mm}
+\psi^-_{\pp}\psi^{\pd}_{\mm})]
\nabla_{\pd}\nabla_{-} {\cal L}|
{}~.
\end{eqnarray}
We start with the following general identity,
derived straightforwardly from
the commutation relations and the Bianchi identities
\begin{eqnarray}
\lefteqn{\nabla^2 \bar{\nabla}^2 +\bar{\nabla}^2\nabla^2 +\nabla_-
\bar{\nabla}^2
\nabla_+
+\nabla_{\md}\nabla^2\nabla_{\pd}} \nonumber\\
&=& \nabla_{\pp}\nabla_{\mm} -\frac{1}{2} \nabla_+R \nabla_- + \frac{1}{2}
\nabla_{\pd}\bar{R}\nabla_{\md}
\nonumber\\
&=&\frac{1}{2} \nabla_{\pp}\nabla_{\mm} +\frac{1}{2} \nabla_{\mm}\nabla_{\pp}
+ {\frac{1}{4}}[- (\nabla_+R)\nabla_- + (\nabla_-R)\nabla_+ +(\nabla_{\pd}
\bar{R})
\nabla_{\md}
-(\nabla_{\md}\bar{R})\nabla_{\pd}]\nonumber\\
&&- \frac{1}{2} [R\nabla^2 -\bar{R}\bar{\nabla}^2] ~~. \label{genident}
\end{eqnarray}
We apply this sum of operators to an arbitrary scalar (so that
certain
connection terms
can be dropped), inside a $d^2x e^{-1}$ integral (so that
space-time
derivatives can
be dropped),
and then evaluate at $\theta =0$, using some of the component
expressions for
covariant
derivatives from Appendix A, etc.
After some lengthy manipulations, that we outline in Appendix C,
we obtain
the following identity:
\begin{eqnarray}
\int d^2 x e^{-1}
\lefteqn{\{\nabla^2 \bar{\nabla}^2 +\bar{\nabla}^2\nabla^2 +\nabla_-
\bar{\nabla}^2\nabla_+
+\nabla_{\md}\nabla^2\nabla_{\pd} \}{\cal L}| } \nonumber\\
&=& \int d^2x e^{-1} \{
[\psi_{\pp}^+ \psi_\mm^- +\psi_{\pp}^- \psi_\mm^+ - \frac{1}{2} B ]
\nabla^2
+[\psi_{\pp}^{\pd}\psi_\mm^{\md}+\psi_{\pp}^{\md}\psi_\mm^{\pd}
+\frac{1}{2}
\bar{B}]\bar{\nabla}^2 \nonumber\\
&+&[\psi_{\pp}^+\psi_\mm^{\md}+\psi_{\pp}^{\md}\psi_\mm^+]
\nabla_+\nabla_{\md}
+[\psi_{\pp}^{\pd}\psi_\mm^- +\psi_{\pp}^-\psi_\mm^{\pd}]
\nabla_{\pd}
\nabla_- \nonumber\\
&+&i\psi_{\pp}^+[\nabla_{\md}\nabla^2 +\nabla_-(\nabla_+\nabla_{\md})]
-i\psi_\mm^-[\nabla_{\pd}\nabla^2 +\nabla_+(\nabla_{\pd}\nabla_-)] \non\\
&+&i\psi_{\pp}^{\pd}[\nabla_- \bar{\nabla}^2 +\nabla_{\md}
(\nabla_{\pd}\nabla_-)]
-i\psi_\mm^{\md}[\nabla_+\bar{\nabla}^2 +\nabla_{\pd}
(\nabla_+\nabla_{\md})] \} {\cal L}|
{}~~.
\end{eqnarray}
On the left hand side of the equation we have the sum of projectors.
We move now certain terms from one side of the equation to the other and
obtain
\begin{eqnarray}
&&\int d^2 x e^{-1}[\nabla^2 -i\psi^{\pd}_{\pp}\nabla_- +i
\psi^{\md}_{\mm}
\nabla_+
-(\psi_{\pp}^{\pd}\psi_\mm^{\md}+\psi_{\pp}^{\md}
\psi_\mm^{\pd} +\frac{1}{2}
\bar{B})]\bar{\nabla}^2 {\cal L}| \nonumber\\
&&+\int d^2 x e^{-1}[\bar{\nabla}^2 -i \psi^+_{\pp}\nabla_{\md} +i
\psi^-_{\mm}\nabla_{\pd}
-(\psi_{\pp}^+ \psi_\mm^- +\psi_{\pp}^- \psi_\mm^+ - \frac{1}{2} B)]
\nabla^2 {\cal L}|
\nonumber\\
&& = \int d^2 x e^{-1}[ \nabla_{\md}\nabla_+ +i \psi^{\pd}_{\pp}
\nabla_{\md}
-i\psi^{-}_{\mm} \nabla_{+} +(\psi^{\pd}_{\pp}\psi^-_{\mm}
+\psi^-_{\pp}\psi^{\pd}_{\mm})]
\nabla_{\pd}\nabla_{-} {\cal L}| \nonumber\\
&&+ \int d^2 x e^{-1} [\nabla_-\nabla_{\pd} + i\psi^+_{\pp} \nabla_{-}- i
\psi^{\md}_{\mm} \nabla_{\pd}
+(\psi^+_{\pp}\psi^{\md}_{\mm} +\psi^{\md}_{\pp}\psi^+_{\mm})]
\nabla_+\nabla_{\md}
{\cal L}|
\end{eqnarray}
However, it can be shown that up to total derivatives the two
terms on each
side of the
equation are equal. We obtain therefore the following result,
which allows us
to
write a full superspace integral in terms of either chiral or twisted
chiral
projections,
\begin{eqnarray}
&& \int d^2 x d^4 \theta E^{-1} {\cal L} \label{twistdensity}\\
&&= \int d^2 x e^{-1} [\nabla^2 -i\psi^{\pd}_{\pp}\nabla_- +i
\psi^{\md}_{\mm}\nabla_+
-(\psi_{\pp}^{\pd}\psi_\mm^{\md}+\psi_{\pp}^{\md}\psi_\mm^{\pd}
+\frac{1}{2}
\bar{B})]\bar{\nabla}^2 {\cal L}|\nonumber\\
&&=\int d^2 x e^{-1}[ \nabla_{\md}\nabla_+ +i \psi^{\pd}_{\pp}
\nabla_{\md}
-i\psi^{-}_{\mm} \nabla_{+} +(\psi^{\pd}_{\pp}\psi^-_{\mm}
+\psi^-_{\pp}\psi^{\pd}_{\mm})]
\nabla_{\pd}\nabla_{-} {\cal L}| ~~.\non
\end{eqnarray}
(The asymmetry between the two forms is due to the constraint
$F=0$.)
Replacing ${\bar{\nabla}}^2 {\cal L}$ or $\nabla_{\pd} \nabla_{-} {\cal L}$ by
arbitrary chiral
or twisted
chiral lagrangians gives the projection formulae for chiral or
twisted chiral
actions,
respectively.
\sect{From full superspace measure to chiral superspace
measure in the
$U_V(1) \otimes U_A(1) $ case}
In this section we derive the analogue of \reff{chint} for the
undegauged
$U_V(1) \otimes
U_A(1)$ supergravity theory. We accomplish this by relating
${\bar{\nabla}}^2$ in
the degauged case to ${{\bar{\nabla}}^2}$ in the $U_V(1) \otimes
U_A(1)$ case.
We consider first the covariant derivative
\begin{equation}
{\nabla_{\pd}} = E_\pd + {\Omega}_\pd \Mb + {\Gamma}_\pd M +
{\S}_\pd {\cal Y} ~~,
\end{equation}
and substitute the explicit expressions for the connections:
\begin{eqnarray}
{\Omega}_\pd &=&+ e^{{S}}(\hat{E}_\md {A_{\pd}^{~\md}} - {A_{\pd}^{~\md} }
\hat{E}_\pd {A_{\md}^{~\pd}} )
\nonumber\\
{\Omega}_\md &=&-e^{{S}}( \hat{E}_\pd {A_{\md}^{~\pd}} - {A_{\md}^{~\pd}}
\hat{E}_\md {A_{\pd}^{~\md}}) \non\\
{\Sigma}_\pd &=& 2ie^{{S}}(\hat{E}_\pd {S}+{A_{\pd}^{~\md}}\hat{E}_\md {S} )
+ie^{{S}}(\hat{E}_\md
{A_{\pd}^{~\md} } -{A_{\pd}^{~\md}}
\hat{E}_\pd {A_{\md}^{~\pd}} ) \non\\
{\Sigma}_\md &=& 2ie^{{S}}(\hat{E}_\md {S} +{A_{\md}^{~\pd}} \hat{E}_\pd {S} )
+ie^{{S}}
(\hat{E}_\pd {A_{\md}^{~\pd}} -{A_{\md}^{~\pd}}
\hat{E}_\md {A_{\pd}^{~\md}} ) \nonumber\\
{\Gamma}_\pd &=& +2e^{{S}}(\hat{E}_\pd {\Sb} + {A_{\pd}^{~\md} }
\hat{E}_\md {\Sb})
+2e^{{\Sb}}
(\hat{E}_\pd {S} +{A_{\pd}^{~\md}} \hat{E}_\md {S}
)+e^{{S}}(\hat{E}_\md {A_{\pd}^{~\md}} -{A_{\pd}^{~\md}} \hat{E}_\pd {A_{\md}^{~\pd}} ) \non\\
{\Gamma}_\md &=& -2e^{{S}}(\hat{E}_\md {\Sb}+{A_{\md}^{~\pd}}
\hat{E}_\pd {\Sb}-2e^{{\Sb}}
(\hat{E}_\md {S} +{A_{\md}^{~\pd}} \hat{E}_\pd {S}
)-e^{{S}}(\hat{E}_\pd {A_{\md}^{~\pd}} -{A_{\md}^{~\pd}} \hat{E}_\md {A_{\pd}^{~\md}}) \non
\end{eqnarray}
We find
\begin{equation}
{\nabla_{\pd} }= E_\pd + {\Omega}_\pd ({\cal M} + i {\cal Y}) +
2E_\pd (S+ \Sb) M
+ 2i(E_\pd S) {\cal Y} ~~. \label{intdelpd}
\end{equation}
We note the following identities
\begin{eqnarray}
e^{-u{\cal Y}} \hat{E}_{\dot{\pm}} e^{u{\cal Y}}
&=& e^{-\frac{i}{2} u} \hat{E}_{\dot{\pm}} + e^ {-\frac{i}{2}u}
(\hat{E}_{\dot{\pm}}u){\cal Y} ~~, \non\\
e^{-vM} \hat{E}_{\dot{\pm}} e^{vM}
&=& \hat{E}_{\dot{\pm}} + (\hat{E}_{\dot{\pm}}v)M ~~.\label{expident}
\end{eqnarray}
We obtain then, from \reff{intdelpd}
\begin{eqnarray}
\lefteqn{e^{2({S}+{\Sb})M +2i {S}{\cal Y}} {\nabla_{\pd}} e^{-2i{S}
{\cal Y} -2({S}
+{\Sb})M}=} \non\\
&&(\hat{E}_\pd + {A_{\pd}^{~\md}} \hat{E}_\md) + (\hat{E}_\md {A_{\pd}^{~\md}} -
{A_{\pd}^{~\md}} \hat{E}_\pd
{A_{\md}^{~\pd})}({\cal M}+i{\cal Y}) ~~. \label{relation}
\end{eqnarray}
We note now that \reff{relation} is valid in both the undegauged
and degauged
theories and the right-hand-side is independent of $S$,
which is the only
quantity that is
affected by the actual degauging.
Therefore the left-hand-sides for the $U_ V(1) \otimes U_A(1)$
and the
$U_A(1)$ theories can be set equal to each other and this
yields a relation
between the covariant
derivatives,
\begin{eqnarray}
\lefteqn{\left( e^{2({\bf S}+\bar{\bf S})M +2i {\bf S}{\cal Y}} {\nabla_{\pd}}
e^{-2i{\bf S}{\cal Y} -2({\bf S}+\bar{\bf S})M}\right)
_{U_V(1) \otimes U_A(1)}
=} \non\\
&&\left( e^{2({S}+{\Sb})M +2i {S}{\cal Y}} {\nabla_{\pd}} e^{-2i{S}
{\cal Y} -2({S}
+{\Sb})M}\right)
_{U_A(1)}
\end{eqnarray}
with a similar expression for $\nabla_{\md}$. Consequently
\begin{equation}
\left({\nabla_\pmd}\right)
_{U_V(1) \otimes U_A(1)}
= e^{2({S}+{\Sb}-{\bf S} - \bar{\bf S})M +2i ({S - {\bf S}})
{\cal Y}} \left(
{\nabla_\pmd}\right)_{U_A(1)}
e^{-2i({S -{\bf S}}){\cal Y} -2({S}+{\Sb} -{\bf S} - \bar{\bf S})M}
~~,
\end{equation}
with a similar relation for the undotted derivatives. To distinguish
between
the two cases, we
have denoted by
${\bf S}$ and $S$ the scale compensators in the $U_V(1) \otimes
U_A(1)$
and the $U_A(1)$
theories, respectively (the former is a general scalar superfield
whereas the
latter is given by
\reff{degauge}).
The above relation expresses the covariant derivatives of the
$U_V(1)
\otimes U_A(1)$ theory
in terms of the covariant derivatives of the $U_A(1)$ theory.
In particular
we can write
\begin{equation}
\left({\bar{\nabla}}^2 \right)_{U_A(1)}
= e^{-2({S}+{\Sb}-{\bf S} - \bar{\bf S})M -2i ({S - {\bf S}}){\cal Y}}
\left({{\bar{\nabla}}^2}
\right)_{U_V(1) \otimes U_A(1)} e^{2i({S -{\bf S}}){\cal Y}
+2({S}+{\Sb}
-{\bf S} - \bar{\bf S})M}~~.
\end{equation}
Acting on a scalar lagrangian ${\cal L}$ we can drop the
exponentials
following ${\bar{\nabla}}^2$, and acting on ${\bar{\nabla}}^2 {\cal L}$
we can drop the
$e^{(\cdots) M}$ term, while
\begin{equation}
e^{-2i( S- {\bf S}){\cal Y}} {\bar{\nabla}}^2_{U \otimes U} {\cal L}
= e^{2(S-{\bf S})} {\bar{\nabla}}^2_{U \otimes U} {\cal L} ~~.
\end{equation}
Substituting this into \reff{chint} we obtain
\begin{eqnarray}
\int d^2x d^4 \theta E^{-1} {\cal L}
&=& \int d^2 x d^2 \theta e^{-2\sigma} (1.e^{\buildrel \leftarrow \over H})\left({\bar{\nabla}^2 }
\right)_
{U_A(1)}{\cal L} |_{\bar{\theta}=0} \non\\
&=& \int d^2 x d^2 \theta E^{-1} \frac{e^{-2{\bf S} }}{[1-{A_{\pd}^{~\md}}
{A_{\pd}^{~\md}}]}
{{\bar{\nabla}}^2}_{U \otimes U} {\cal L} |_{\bar{\theta}=0}
~~, \label{undegaugedchint}
\end{eqnarray}
which defines the chiral measure in the $U_V(1) \otimes U_A(1)$
theory,
and allows us to
construct invariant actions of the form $\int d^2x d^2 \theta {\cal E}^{-1}
{\cal L}_{chiral}$.
Obviously, since the theory is symmetric under the
interchange of $-$
and $\md$, a twisted
chiral measure exists for writing invariants $\int d^2x d^2 \tilde{\theta}
\tilde{{\cal E}}^{-1} {\cal L}_{twisted~chiral}$. However,
as we discuss
in the concluding section, an explicit
expression is harder to come by.
\sect{Discussion}
In this paper we have presented results for the integration
measures in the
$(2,2)$ theories,
and have derived projection formulae for obtaining component
actions from the
corresponding superspace actions. This particular issue has been
a sore point in
most derivations because no clear, completely superspace
technique
exists in the general case. For the case at hand, at least, we have
managed
to avoid
any reference to component supersymmetry transformations
\cite{Superspace}
or explicit $\theta$ expansions of the supergravity prepotentials
in Wess-Zumino
gauge
\cite{Buchbinder}.
Whereas in the degauged theory, which contains a chiral
compensator, the
existence and
construction of the chiral measure is straightforward, this is
not obviously
the case in
the undegauged theory. We have shown, by explicitly
expressing the covariant
derivatives
of the $U_V(1) \otimes U_A(1)$ theory in terms of those
of the degauged,
$U_A(1)$ theory,
how to obtain the chiral measure for the former theory,
from a knowledge of
the chiral measure for the latter theory.
We have also argued that, by symmetry, a twisted chiral
measure exists. However,
its explicit construction is not straightforward. The point is
that in our
solution of the
constraints in ref. \cite{MGMW}, we have chosen to express
the covariant
derivatives in terms of the
objects $\hat{E}$ and the compensators $S$ which break the
symmetry between
$-$ and $\md$ objects. To maintain the symmetry we would
have to introduce
additional prepotentials which were eliminated from the
beginning by an
appropriate $K$-transformation gauge choice.
The lack of symmetry manifests itself also in the construction
of covariantly
chiral
and covariantly twisted chiral superfields in terms of ordinary
chiral and
twisted
chiral ones. The construction of covariantly chiral superfields is
straightforward,
\begin{equation}
\Phi_{cov.~chiral} = e^H \phi_{chiral} \label{covchi}~~,
\end{equation}
because then the left-hand-side is annihilated by
$\nabla_{\dot{\pm}}$ when the
right-hand-side is annihilated by $D_{\dot{\pm}}$ (c.f. (2.7)).
However, if we want
to obtain a
covariantly twisted chiral superfield ${\cal X}$ satisfying
\begin{equation}
\nabla_{\pd} {\cal X} = \nabla_- {\cal X} =0
\end{equation}
in terms of an ordinary twisted chiral superfield satisfying
\begin{equation}
D_{\pd} \chi = D_- \chi =0 ~~,
\end{equation}
it is clear that the above construction will not work. In fact,
we have not
succeeded
in obtaining a closed form relation similar to \reff{covchi}. In the
remainder of this
concluding section we show what the relation is to first order in the
supergravity
prepotential.
We begin by writing down the differential equations that the twisted
chirality conditions imply for ${\cal X}$,
\begin{eqnarray}
{[} e^{-H}D_-e^H + A_-^{~+} e^{-H} D_+ e^H ] {\cal X} &=& 0 \non \\
{[} e^H D_\pd e^{-H} + A_{\pd}^{~\md} e^H D_\md e^{-H}] {\cal X}
&=& 0 ~~.
\end{eqnarray}
To first order in $H$, these equations reduce to
\begin{eqnarray}
D_- {\cal X} + i (D_-H^a) \partial_a {\cal X} + A_-^{~+} D_+ {\cal X} &=& 0 \non \\
D_\pd {\cal X} - i (D_\pd H^a) \partial_a {\cal X} + A_{\pd}^{~\md} D_\md {\cal X}
&=& 0 ~~,
\end{eqnarray}
with
\begin{eqnarray}
A_-^{~+} &=& -2 D_\pd D_- H^\pp + {\cal O}(H^2) \non\\
A_{\pd}^{~\md} &=& -2 D_\pd D_- H^\mm + {\cal O}(H^2)~~.
\end{eqnarray}
We now set ${\cal X} = \chi + Z$, and solve iteratively for $Z$.
To linear
order we find
that
\begin{equation}
Z = -2 D_\pd H^\pp D_+ \chi + 2 D_- H^\mm D_\md \chi +
i H^\mm \partial_\mm \chi
- i H^\pp \partial_\pp \chi ~~,
\end{equation}
and therefore we can express ${\cal X}$ and, in a similar fashion,
$\bar{{\cal X}}$ as
\begin{eqnarray}
{\cal X} &=& [1-2 D_\pd H^\pp D_+ + 2 D_- H^\mm D_\md +
i H^\mm \partial_\mm
- i H^\pp \partial_\pp] \chi + {\cal O}(H^2) \non \\
\bar{{\cal X}} &=& [1+2 D_+ H^\pp D_\pd - 2 D_\md H^\mm D_- -
i H^\mm \partial_\mm + i H^\pp \partial_\pp] \bar{\chi} + {\cal O}(H^2) ~~.
\end{eqnarray}
{\bf Acknowledgments}
The impetus for much of this work came from our interaction
with Jim Gates.
We wish to thank him for discussions, suggestions, and
general insights into many aspects of superspace and its
geometry.
We also thank Nathan Berkovits for discussions concerning
the existence of
chiral and twisted chiral measures in the $U_V(1) \otimes
U_A(1)$ theory.
M.E.W. thanks the Physics Department of Queen's University
for hospitality.
\newpage
|
\section*{Introduction}
Gamma-ray bursts (GRBs) continue to confound astrophysicists nearly a
quarter century after their discovery \cite{KSO:739}. Before the
launch of CGRO, most scientists thought that GRBs came from magnetic
neutron stars residing in a thick disk (having a scale height of up to
$\sim$ 2 kpc) in the Milky Way \cite{Hig:Lin:90,Hard:91}. The data
gathered by BATSE showed the existence of a rollover in the cumulative
brightness distribution of GRBs and that the sky distribution of even
faint GRBs is consistent with isotropy \cite{Meegan:92,Briggs:95}.
This rules out a thick Galactic disk source population.
Consequently, the primary impact of the BATSE results has been to
intensify debate about whether the bursts are Galactic or cosmological
in origin. Galactic models attribute the bursts primarily to
high-velocity neutron stars in an extended Galactic halo, which must
reach one fourth or more of the distance to M31 ($d_{\rm M31} \sim 690$
kpc) in order to avoid any discernible anisotropy
\cite{Hak:94,Hartmann:94}. Cosmological models place the GRB sources
at distances $d \sim 1 - 3$ Gpc, corresponding to redshifts $z\sim 0.3
- 1$. A source population at such large distances naturally produces
an isotropic distribution of bursts on the sky, and the expansion of
the universe or source evolution can reproduce the observed rollover in
the cumulative brightness distribution \cite{Fenimore:93}.
Recent studies \cite{LyneLori:94,Frail:94} have revolutionized our
understanding of the birth velocities of radio pulsars. They show that
a substantial fraction of neutron stars have velocities that are high
enough to produce an extended halo around the Milky Way like that
required by Galactic halo models of GRBs \cite{LiDer:92}.
Podsiadlowski, Rees, and Ruderman \cite{Pods:94} have carried out
pioneering calculations of the spatial distribution expected for
high-velocity neutron stars born in the Galactic disk. They consider
the effects of a non-spherical halo potential and of M31, but neglect
the effects of the Galactic disk, which we find is also important.
\section*{Models}
We have calculated detailed models of the spatial distribution expected
for a population of high-velocity neutron stars born in the Galactic
disk and moving in a Galactic potential that includes the bulge, disk,
and a dark matter halo.
We use the mass distribution and potential given by Kuijken and Gilmore
\cite{KG:89} which includes a disk, a bulge, and a dark matter halo.
The densities of the disk and of the halo are
\begin{eqnarray}
\rho_D= \rho_D^0 \exp\left({-r\over r_d}\right)\exp\left({-z\over z_d}\right),
&
{}~~~~~~ &
\rho_H=\rho_H^0 \left[ 1 + \left({r\over r_c}\right)^2 \right]^{-1}.
\end{eqnarray}
The circular velocity $v_c$ and the Galactic disk lead to
characteristic angular anisotropies as a function of burst brightness
which provide a signature, and therefore a test, of high-velocity
neutron star models. Prolate or oblate dark matter halos also produce
other angular anisotropies as a function of burst brightness which may
provide a signature of such models \cite{Pods:94}.
We assume that neutron stars are born with the circular velocity $v_c
\approx 220~\hbox{km~s$^{-1}$}$ of the Galactic disk. Given that current knowledge
of the distribution of initial kick velocities is uncertain, we adopt a
Green-function approach: we calculate the spatial distribution of
neutron stars for a set of kick velocities (e.g., $v_{\rm kick} = 200,
400,..., 1400$~\hbox{km~s$^{-1}$}). We follow the resulting orbits for up to $3
\times 10^9$ years.
In our initial calculations, we assume that the bursts are standard
candles, i.e. $L = \delta(L-L_0)$. We parameterize the burst-active
phase by a turn-on age $\delta t$ and a duration $\Delta t$, and assume
that the rate of bursting is constant throughout the burst-active
phase. The high-velocity neutron star model then has four parameters:
$v_{\rm kick}$, $\delta t$, $\Delta t$, and the BATSE sampling depth
$d_{\rm max}$.
\begin{figure}[th]
\begin{center}
\begin{tabular}{lr}
{\psfig{file=plo2cth.ps,width=5.cm,angle=-90}} &
{\psfig{file=plo2sb2.ps,width=5.cm,angle=-90}} \\
{\psfig{file=plo2pf2.ps,width=5.cm,angle=-90}} &
{\psfig{file=brightD.ps,width=5.cm,angle=-90}} \\
\end{tabular}
\end{center}
\caption{Comparison of a Galactic halo model in which neutron stars
are born with a kick velocity of
$1000$~\hbox{km~s$^{-1}$}\ and have a burst-active phase lasting
$\Delta t = 500$ million years with a carefully-selected sample of 285
bursts from the BATSE 2B catalogue. Panels (a) and (b) show the
contours in the ($\delta t$, $d_{\rm max}$)-plane along which the
Galactic dipole and quadrupole moments of the model differ from those
of the data by $\pm$ 1$\sigma$ (solid lines), $\pm$ 2$\sigma$ (dashed
line), and $\pm$ 3$\sigma$ (short-dashed line) where $\sigma$ is the
model variance; the thin line in panel (a) shows the contour where the
dipole moment for the model equals that for the data. Panel (c) shows
the contours in the ($\delta t$, $d_{\rm max}$)-plane along which 32\%,
5\%, and $4 \times 10^{-3}$ of simulations of the cumulative
distribution of 285 bursts drawn from the peak flux distribution of the
model have KS deviations $D$ larger than that of the data. Panel (d)
compares the brightness distribution of the model shown in (a) - (c),
taking $\delta t=30$~Myrs and $d_{max}=200$~kpc, to the BATSE plus PVO
data.
}
\vspace{-5mm}
\end{figure}
\section*{Comparison between models and data}
We compare the models with a carefully-selected data set that is
self-consistent. We use only bursts that trigger on the 1024~ms
timescale because we require that all bursts lie above the counts
threshold in one trigger timescale; the 1024~ms timescale yields the
largest sampling depth, and therefore imposes the strongest constraint
on models, of the three BATSE trigger timescales. We adopt $F_{\rm
pk}^{1024}$, the peak flux in 1024~ms, as our measure of burst
brightness. We therefore include only bursts which have a $F_{\rm
pk}^{1024}$ and $t_{90} > 1024$~ms. We consider only bursts with
$F_{\rm pk}^{1024} \ge 0.35$~photons~cm$^{-2}$~s$^{-1}$ in order to
avoid threshold effects \cite{Fenimore:93,ZandtFen:94}. We also
exclude overwriting bursts, because the threshold is much higher for
these bursts, and MAXBC bursts, because they have unknown positional
errors. The 2B catalogue contains 285 bursts satisfying the above
criteria. This set of bursts has Galactic dipole and quadrupole
moments $\langle \cos \theta \rangle =0.056 \pm 0.034$, and $\langle
\sin^2 b-{1\over 3} \rangle = -0.033 \pm 0.017$, compared to the values
$\langle \cos \theta \rangle =-0.013$, and $\langle \sin^2 b-{1\over 3}
\rangle = -0.005$ expected for a uniform sky distribution, taking into
account the BATSE sky exposure.
As a first step in testing the viability of Galactic halo models, we
have compared the Galactic dipole and quadrupole moments, $\langle \cos
\theta \rangle$ and $\langle \sin^2 b - 1/3 \rangle$, of the angular
distribution of bursts for the model with those for the above set of
bursts, using $\chi^2$. We have also compared the peak flux
distribution for the model with that for the above set of bursts, using
the KS test.
\begin{figure}[t]
\centerline{\psfig{file=map.285.ps,width=11cm,angle=-90}}
\caption{Sky distribution of 285 bursts drawn randomly
from the angular distribution expected for the model
illustrated in Figure~1 when $\delta t=30$~Myrs, and
$d_{\rm max}= 200$~kpc.}
\end{figure}
These comparisons do {\it not} provide estimates of model parameters
(i.e., they do not yield parameter confidence regions), but are meant
only to be a rough ``goodness-of-fit" guide to models which should be
tested using a more rigorous approach like the maximum likelihood
method.
As an illustrative example, we show in Figure~1 the results for a
Galactic halo model in which neutron stars are born with a uniform
single velocity $1000$~\hbox{km~s$^{-1}$}\ and the burst-active phase has an initial
burst rate $r \propto t^2$ and lasting $\Delta t = 500$ million years.
Figure 2 shows the sky distribution of 285 bursts drawn randomly from
the angular distribution expected for the model with $\delta
t=30$~Myrs, and $d_{\rm max}= 200$~kpc.
Comparisons of this kind show that the high-velocity neutron star model
can reproduce the peak flux and angular distributions of the bursts in
the BATSE 2B catalogue for neutron star kick velocities $v_{\rm kick}
\mathrel{\mathpalette\simov >} 800$~\hbox{km~s$^{-1}$}, burst turn-on ages $\delta t \mathrel{\mathpalette\simov >}
10$~million years, and BATSE sampling depths $100$~kpc $\mathrel{\mathpalette\simov <} d_{\rm
max} \mathrel{\mathpalette\simov <} 400$~kpc. Moreover, comparisons of this kind show that
there is a large region of parameter space in which these models can
reproduce the angular distribution of the bursts in the preliminary
BATSE~3B catalogue.
In high-velocity neutron star models, the slope of the cumulative peak
flux distribution for the brightest BATSE bursts and the PVO bursts
reflects the space density of the relatively small fraction of burst
sources in the solar neighborhood. The nearness of the observed slope
of the cumulative peak flux distribution of these bursts to -3/2, the
value expected for a uniform spatial distribution of sources which emit
bursts that are ``standard candles," must be considered a coincidence
in the high-velocity neutron star model. However, a spread in neutron
star kick velocities, in neutron star ages at which bursting behavior
begins, or in the burst luminosity function tends to produce a
cumulative peak flux distribution with a slope of -3/2; beaming of
bursts along the direction of motion of the source or evolution of the
rate of bursting as a function of age also tends to produce a slope of
-3/2.
We find that there are many combinations of these factors which
successfully reproduce the slope of the BATSE plus PVO peak flux
distribution. For example, a model in which the burst luminosity
function is a log normal distribution with a FWHM of a factor of
$\mathrel{\mathpalette\simov <} 10$ and the burst-active phase has an abrupt (``heaviside
function") turn-on, one in which the kick velocities are distributed
between $800$ and $ 1200$~km~s$^{-1}$ and the burst-active phase
initially has a burst rate $r = (t/\delta t)$, or one in which the
burst-active phase initially has a burst rate $r = {1 \over 2}
(t/\delta t)^2$ work equally well. Figure~1d compares the peak flux
distribution for the last model and the BATSE+PVO peak flux
distribution.
M31 provides a strong constraint on the BATSE sampling distance $d_{\rm
max}$ \cite{Hak:94}. We have investigated the effects of M31 within the
framework of the high-velocity neutron star model described above by
including the distortion of the Galactic halo potential due to M31 and
the burst sources emanating from M31. We find that for such models M31
imposes a limit on the BATSE sampling distance $d_{\rm max} \mathrel{\mathpalette\simov <}
400$~kpc, even if the bursting activity of neutron stars lasts for more
than $10^9$ years \cite{BulCopLam:95c}.
|
\section{Abstract}
Gravitational lensing is one of a number of methods
used to probe the distribution of dark mass in the Universe.
On galactic scales, complementary techniques include the use of
stellar kinematics, kinematics and morphology of the neutral gas layer,
kinematics of satellites, and morphology and temperature profile
of X-ray halos. These methods are compared, with emphasis on their
relative strengths and weaknesses in constraining
the distribution and extent of dark matter in the Milky Way and
other galaxies. It is concluded that (1) the extent of dark halos
remains ill-constrained, (2) halos need not be isothermal,
and (3) the dark mass is probably quite flattened.
\section{Introduction}
Modeling the gravitational structure of a galaxy, and therefore its
lensing properties, requires knowledge of the extent,
radial profile, and geometric form of its mass distribution.
The interpretation of microlensing rates and optical
depths along different lines of sight through the Milky Way,
for example, is strongly dependent on the assumed distribution of
total (light and dark) Galactic mass. On larger
scales, efficient and reliable image-inversion techniques designed
to measure the structure parameters of intervening lensing galaxies
require appropriate fitting functions for the lensing mass.
This review focuses on techniques that
form a symbiotic relationship with lensing in producing
valuable and complementary constraints on galactic potentials,
especially in providing partial answers to the
following questions about galactic {\it dark\/} mass:
\begin{itemize}
\item What is the physical extent of dark matter in galaxies?
\item Is the distribution of dark mass isothermal?
\item What is the shape of dark ``halos''?
\end{itemize}
\section{How Big are Dark Halos?}
The size of dark halos controls the galactic ``sphere of influence''
for lensing, interactions, and accretion.
Together with the radial and vertical structure parameters,
halo extent determines the total mass of the galaxy.
The notion of halos as distinct entities ceases
to be useful, of course, on scales larger than half the mean
distance to the nearest, comparably-sized neighbor.
If halos are extremely large
and isothermal, they cannot be totally baryonic without violating
the constraints on $\Omega_B h^2$
from primordial big bang nucleosynthesis (BBN) models. Recent
assessments give $0.01 \leq \Omega_B \leq 0.06$ for $0.5 \leq h_{100} \leq 1$
(Walker {\it et~al.}\ 1991, Smith, Kawano \& Malaney 1993).
Since the density of observed baryons is $\Omega_{\rm Lum} \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.007$
(Pagel 1990), the average ratio of dark-to-luminous baryons is
$0.4 \leq M_{B,\rm Dark}/M_{B,\rm Lum} \leq 8$,
and at least some of the Universe's baryons are dark.
Rotation curve analysis indicates that
$1 \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} M_{B,\rm Dark}/M_{B,\rm Lum} \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 10$ ({cf.}\ Broeils 1992),
so that on scales comparable to {H{\sc I}}\ disks ($\sim$30~kpc),
halos composed entirely
of dark baryons are consistent with BBN.
Faint galaxies are more numerous and more dark matter dominated
than brighter galaxies, but contribute less to the
total luminosity of the Universe.
Thus the upper limit placed by BBN on the size of
baryonic halos is likely to be considerably larger than 30~kpc
(Binney \& Tremaine 1987), but its
calculation requires a model-dependent integral over the
galaxy luminosity function, weighted by $M_{B,\rm Dark}/M_{B,\rm Lum}$.
\subsection{The Extent and Mass of the Milky Way Halo}
At large radius, the mass of our galaxy can be estimated
from the kinematics of distant, presumably bound, objects --- halos
stars, satellite galaxies, and group members ---
and from the kinematics of
Magellanic Clouds/Stream system. The former has been done most
recently by Kochanek (1995), who finds that, using a Jaffe model
as the global mass distribution for the Galaxy,
the total mass inside 50~kpc at 90\% confidence
is $(5.4 \pm 1.3) \times 10^{11} \rm {M_{\odot}} $
if the timing constraints of the Local Group are imposed
and Leo I is bound, and somewhat lower at
$4.3^{+1.8}_{-1.0} \times 10^{11} \rm {M_{\odot}} $
if the timing constraints are not imposed.
The corresponding masses within 100~kpc are
$7^{+4}_{-3} \times 10^{11} \rm {M_{\odot}} $ and
$(8 \pm 2) \times 10^{11} \rm {M_{\odot}} $, respectively.
A recent re-examination of the kinematics and proper motions of
the Magellanic Clouds and Stream using two different
model potentials for the Milky Way (Lin, Jones \& Kremola 1995)
yields $(5.5 \pm 1) \times 10^{11} \rm {M_{\odot}} $ inside
100~kpc. About one-half this mass must lie {\it outside\/} the
present Cloud distance (50~kpc) in order to explain with
these models the observed infall of the Magellanic Stream.
Since the luminous matter in the Galaxy accounts
for $(0.6-1) \times 10^{11} \rm {M_{\odot}} $,
the full range of dark mass estimates from these two methods is
$1.3 \times 10^{11} \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} M_{\rm Dark}(<50 \rm kpc) \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}}
6.1 \times 10^{11} \rm {M_{\odot}} $,
with apparent contradictions at the lower end with
Local Group timing and at the upper end with Magellanic
Stream kinematics.
The implied upper limit of $M_{\rm Dark}(<50 \, \rm kpc) \sim2.7
\times 10^{11} \rm {M_{\odot}} $ from the Magellanic Stream model is
only just consistent with the lower limit of $\sim2.3 \times 10^{11} \rm {M_{\odot}} $
from the satellite model.
For comparison, the spherical isothermal
dark halo used by many microlensing teams as a fiducial model
contains $4.1 \times 10^{11} \rm {M_{\odot}} $
interior to the LMC distance of 50 kpc (Griest 1991).
Using this model and its first year's LMC data, the MACHO team
concludes with 68\% confidence that the
total mass in compact dark lenses is $7^{+6}_{-4} \times 10^{10} \rm {M_{\odot}} $
(Alcock {\it et~al.}\ 1995).
(For more complete reviews of the mass of the Galaxy and its dependence
on assumptions about Leo I, see Fich \& Tremaine 1991, Schechter 1993, and
Freeman, these proceedings.)
\subsection{Halo Size of External Galaxies}
The kinematics of satellites can also be used to study
the halos of external galaxies, but since only a small number of
satellites are observed per primary,
conclusions are based on a statistical analysis of the
sample as a whole. Based on satellite velocities and {H{\sc I}}\ rotation curves,
Erickson, Gottesman and Hunter (1987) concluded that
the primaries in their sample have $M_{Dark}/M_{Lum} < 5$,
total $M/L \sim 20$, and potentials that
are well-described by a point mass model --- all consistent
with dark halos that extend no more than 3 disk radii.
In a more recent study using a different sample, however,
Zaritsky and White (1994) conclude that halos are nearly isothermal, with
total $M(<200 \rm kpc) = 1.5-2.6 \times 10^{12} \rm {M_{\odot}} $
and $110 < M/L < 340$ (for $h_{100}=3/4$).
Their result is primarily due to secondaries at 200-300~kpc,
where the orbital times are on the order of a Hubble time, thus
necessitating the use of halo formation models
to interpret the satellite kinematics.
Using their method, Zaritsky and White conclude that
the Erickson {\it et~al.}\ sample,
which has smaller mean primary-satellite separation,
is consistent with both small and large mass halos.
In the future, weak lensing is likely to play a larger
role in constraining the extent of dark halos. Recent work by
Brainerd, Blandford and Smail (these proceedings)
has given the first indication that the tangential distortion of background
galaxies due to weak lensing by foreground galaxies is statistically
measurable for a large sample ($\sim 3000$) of source-lens pairs.
Their measurement
of $1.0^{+1.1}_{-0.7} \times 10^{12} h^{-1} \rm {M_{\odot}} $
for the total mass within 100 $h^{-1}$ kpc is consistent both
with a mass distribution that grows linearly to 100~kpc
and one that truncates much sooner with total $M/L \approx 10$.
The ring of {H{\sc I}}\ gas in the M96 group in Leo (Schneider 1985) offers
the rare opportunity to sample galactic potentials at very large radii
using the well-defined orbits of cold gas. The Leo ring has a
radius of 100 kpc and completely encircles the early-type
galaxies M105 and NGC~3384.
The radial velocities and spatial distribution
of the gas are consistent with a single, elliptical {\it Keplerian\/}
orbit with a center-of-mass velocity equal to the centroid of
the galaxy pair, and a focus that can be placed at the barycenter of
the system without compromising the fit. The implied
dynamical mass within 100~kpc is $5.6 \times 10^{11} \rm {M_{\odot}} $,
(only twice that inferred from the internal dynamics of the galaxies),
giving a total $M/L \approx 25$.
The sensitivity of non-circular orbits
to the power law form of the potential suggests that dark
matter does not extend much beyond the ring pericenter
radius of 60~kpc.
As a caveat, it is yet clear to what degree M96, a spiral
located 60~kpc (in projection) outside the ring, may perturb
the ring kinematics.
\section{Are Dark Halos Isothermal?}
The approximate flatness of {H{\sc I}}\ rotation curves
is the best observational evidence that dark matter is
present in spirals and has a shallower density profile than
the light; an isothermal halo is
as shallow as $r^{-2}$. Early theoretical studies (Gott 1975)
suggested that violent relaxation would cause the inner regions of
galaxies to have steeply falling profiles that
would flatten to $r^{-2.25}$ in the outer parts.
More recent CDM models (Navarro, Frenk \& White 1995)
indicate that dark halo profiles may be shallower than $r^{-2}$ in
the center and quite steep near the virial radius.
Compression by a dissipating gaseous disk
may further contract and flatten
the dark matter (Blumenthal {\it et~al.}\ 1986), accounting in part for
the apparent ``conspiracy'' between the
dark and luminous mass that produces flat rotation curves.
\subsection{Milky Way}
The radial structure of the mass and light in
the Milky Way is less well-constrained than in external
galaxies. Determining the rotation curve of the Galaxy, in particular,
has proven notoriously difficult.
On the other hand, distances and kinematics of old, resolved stars
can be used to measure the vertical restoring force
of the local disk --- and thus its surface mass density.
In this way, Kuijken and Gilmore (1991) report a mass column of
$71 \pm 6 \rm {M_{\odot}} pc^{-2}$ within a 1.1~kpc band from the
Galactic plane,
with $48 \pm 9 \rm {M_{\odot}} pc^{-2}$ due to the disk itself, and
the rest contributed by a rounder halo. Other recent estimates
are similar: Gould (1990) weighs in at $54 \pm 8 \rm {M_{\odot}} pc^{-2}$,
Bahcall, Flynn and Gould (1992) at $54 \pm 8 \rm {M_{\odot}} pc^{-2}$, and
Flynn and Fuchs (1994) at $52 \pm 13 \rm {M_{\odot}} pc^{-2}$.
The dynamical disk mass thus seems to be in
remarkable agreement with the detectable disk mass of
$49 \pm 9 \rm {M_{\odot}} pc^{-2}$ --- at least locally,
almost none of the disk mass is dark.
Since only about one-half of the local rotation support
is provided by the observable disk, this further implies that
dark matter in the Galaxy is dynamically important
at radii as small as 2.5 disk scale lengths.
Stated in the language of \S4.2,
the Milky Way disk is one-half of its ``maximal disk'' value.
Unfortunately, uncertainties in the outer Galactic rotation
curve frustrate attempts to determine the distribution of mass in
the outer Galaxy, which is further complicated by a recent suggestion that the
generally-accepted local rotation speed, $\Theta_0 = 220$ {\rm km s$^{\hbox{\tiny --1}}$},
may be overestimated by $\sim$10\% (Merrifield 1992).
A smaller value would increase the relative dynamical importance
of the luminous disk and decrease the slope of the outer
rotation curve, to which $\Theta_0$ is tied.
Conclusions drawn from microlensing results about the
dark baryonic content of the Milky Way
depend on the assumed distribution
of dark {\it and\/} luminous matter in the Galaxy
({cf.}\ Paczy\'nski, these proceedings).
Many studies have explored how different assumptions for $M/L$,
rotation curve slope, and the shape, truncation radius and
radial profile of the halo affect these
conclusions ({cf.}\ references in Griest {\it et~al.}\ 1995).
As an indication of the importance of {\it luminous\/} structure,
lensing by stellar bars in the Milky Way and the LMC
has been held accountable, respectively, for most of the optical depth toward
the Galactic center (Zhao, Spergel \& Rich 1995) and the
LMC (Sahu 1994). On the other hand, if the Galactic disk were ``maximal, ''
the MACHO results toward the LMC would be consistent with a dark halo
entirely composed of lensing baryons
(Alcock {\it et~al.}\ 1995).
\subsection{Radial Distribution of Dark Mass in External Galaxies}
In contrast to the difficulties in the Milky Way,
surface brightness profiles and rotation curves for external galaxies
can be measured well,
but their disk mass-to-light ratios, $M/L$, are uncertain.
A disk $M/L$ that is constant with radius (but varies from galaxy to
galaxy) can explain the kinematics within the optical radius of
many spirals ({cf.}\ Kalnajs 1983, Kent 1986, Buchhorn 1992), but
the high velocities observed at the edges of {H{\sc I}}\ disks can be
reproduced only by invoking a rapid radial increase in $M/L$
({cf.}\ Kent 1987, Begeman 1987).
Since the age and metallicity gradients inferred from the
blueing radial color gradients in spirals do not
produce these strong, {\it positive\/} gradients in $M/L$
({cf.}\ de~Jong 1995), dark matter is implicated.
In order to estimate conservatively the amount of
dark matter in a galaxy, the ``maximum disk hypothesis'' is often adopted
(van Albada \& Sancisi 1987), which fixes the disk $M/L$
at the value that maximizes the disk mass
without violating kinematic constraints.
The hypothesis is controversial ({cf.}\ Rubin 1987,
Casertano \& van Albada 1990, Freeman 1993),
but when it is used to fit rotation curves, the resulting disk $M/L$
are larger for brighter and earlier type spirals than for
fainter and later type spirals (Broeils 1992, Buchhorn 1992).
The correlation appears to be stronger in bluer bands.
These trends may be due to the older stellar populations
associated with early spirals,
a notion supported by comparison with the $M/L$ derived
from stellar population synthesis models
(Athanassoula, Bosma \& Papaioannou 1987) and the observed
stellar dispersions in spirals (van der Kruit \& Freeman 1986).
Alternatively, they may reflect trends in dark matter properties
with galaxy type and luminosity that are incorrectly characterized by
the application of maximum disk models (van der Kruit 1995).
\begin{figure}
\vskip -1.3cm
\epsfxsize=\hsize\epsffile{SackettIAU173Fig1.ps}
\vskip -2.0cm
\caption{
Inversion of rotation curves (left)
to derive dynamical surface mass densities, $\Sigma$ (right).
Two extrapolations for $v(r)$ are shown: flat
(dotted) and Keplerian (dashed).
{\it Top: Thin, exponential disk.\/} Both extrapolations overestimate
the true $\Sigma$ (solid) because for an
exponential disk $v(r)$ declines faster than Keplerian.
Only for inner half of the disk can $\Sigma$ be determined reliably.
A spherically-symmetric mass estimator is unreliable for an exponential
disk and can produce negative $\Sigma$ in the inversion.
{\it Bottom: Sbc NGC~2903.\/} Real data (solid squares) and
artificially noisy data (open squares) are
inverted using both extrapolation schemes.
$M/L \, $ increases markedly beyond $\sim$2 scale lengths.
Adding noise makes little difference. A spherical isothermal
halo has a $\Sigma$ that is $\sim\pi/2$ larger than that
of the flat extrapolation, but with similar slope (Sackett, in preparation).
}
\vskip -0.5cm
\end{figure}
It should be stressed that {\it rotation curves do not
constrain the dark matter
distribution to have a $r^{-2}$ (isothermal) volume density profile\/}.
If (1) rotation curves were perfectly flat, (2) halos were spherical,
and (3) the luminous mass were negligible,
then indeed dark matter halos could be described by
singular isothermal spheres over the radial range of the kinematics.
In fact, rotation curves are seldom flat,
but instead have slopes that are
systematically related to the peak speed or the luminosity
concentration of the galaxy (Kent 1987, Athanassoula, Bosma \&
Papaioannou 1987, Casertano \& van Gorkom 1991, Broeils 1992):
diffuse, slow rotators have rising rotation curves,
while compact, fast rotators have falling curves.
Evidence is mounting that dark halos are not spherical (\S 5).
Finally, since the stellar mass is not strongly constrained,
dark halos with asymptotic $r^{-3}$ and $r^{-4}$ density profiles
are also consistent with observed rotation curves (Lake \& Feinswog 1989).
Even when $r^{-2}$ halos are used to fit rotation curves ---
together with luminous disks of reasonable $M/L$ ---
a large core radius must be assumed, so that the halo does not
achieve its asymptotic (isothermal) speed at the last measured point
(Fig.~1). This is especially true of maximum disk fits
({cf.}\ Broeils 1992), but often applies to fits that assume
smaller disk masses as well ({cf.}\ Kent 1987).
This suggests that the linewidth of the {H{\sc I}}\ gas used
in the Tully-Fisher relation is probably not governed
by the asymptotic speed of an isothermal dark halo.
Rotation curve inversion is a step toward a model-independent
method for determining the radial distribution of dark mass in galaxies.
The technique has been criticized as being sensitive to noise (Binney \&
Tremaine 1987), but for this application the typical uncertainties of 10-20\%
are quite tolerable.
The method does depend on the assumed geometry
of the mass and extrapolation of the rotation curve beyond the
last measured point (Fig.~1), but has the advantage of making
this dependence explicit rather than camouflaging it by the use of
a particular model for the dark mass.
\section{What is the Shape of the Dark Mass: Disks or Halos?}
Use of term ``halo'' to describe
the distribution of dark matter may be prejudicial: there is
no strong theoretical or observational evidence to indicate that
dark matter in galaxies is distributed spherically.
Dark halos have been favored over dark disks as a means
to stabilize galaxies against bar formation (Ostriker \& Peebles 1973),
but bulges (Kalnajs 1987) and hot disks (Athanassoula \& Sellwood 1987)
are now believed to be more efficient stabilizers.
Traditional rotation curve analysis is insensitive to vertical
structure, but accumulating observational evidence from other
methods suggests that the dark mass may be considerably flattened toward the
stellar plane, while remaining relatively axisymmetric,
with an in-plane axis ratio of $(b/a)_\rho > 0.7$
(see review by Rix 1995).
Here, we focus on $(c/a)_\rho$, the vertical flattening of
dark matter, since it is likely to have the stronger implications for
both microlensing in the Galaxy
({cf.}\ Gould, Miralda-Escud\'e, \& Bahcall 1994),
and the use of macrolensing as a probe of galaxy structure.
The flattening of the dark mass may also provide a clue
as to the nature of its constituents.
N-body simulations of dissipationless collapse produce strongly
triaxial dark halos (Frenk {\it et~al.}\ 1988,
Dubinski \& Carlberg 1991, Warren {\it et~al.}\ 1992), but adding a
small fraction ($\sim$10\%) of dissipative gas results in halos
of a more consistent shape --- nearly oblate, $(b/a)_\rho \mathrel{\raise.3ex\hbox{$>$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.8$,
but moderately flattened, $(c/a)_\rho \mathrel{\raise.3ex\hbox{$>$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.6$
(Katz \& Gunn 1991, Dubinski 1994).
Thus strongly flattened halos, with $(c/a)_\rho < 0.5$, may imply that
dissipation has played an even greater role, perhaps implicating
baryonic dark matter.
In order to measure $(c/a)_\rho$ of
the dark mass, a probe of the vertical gradient of the potential
is required.
In the Milky Way, the measured anisotropy of the velocity dispersion
of extreme Population II halo stars has been used to estimate
the flattening of the mass distribution (Binney, May \& Ostriker 1987,
van~der~Marel 1991). Unfortunately, the results depend on the
unknown orbital structure of the stellar halo, so that $(c/a)_\rho$ can
be confined only to lie between 0.3 and 1 at the solar
neighborhood.
In external galaxies, Buote and Canizares (1994, 1995) have used
the flattening of extended X-ray isophotes, assuming
that the gas is in hydrostatic equilibrium, to place constraints on
the flattening of the dark matter in two early-type systems.
For the elliptical NGC~720, they find that the dark isodensity contours
have axis ratio $0.3 \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} (c/a)_\rho \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.5$ at 90\% confidence;
for the lenticular NGC~1332, $0.2 \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} (c/a)_\rho \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.7$.
This suggests that these dark halos are at least as flattened as
their corresponding luminous galaxies, which have optical isophotes
of axis ratio $q \approx 0.6$.
These values contrast with that of $(c/a)_\rho \geq 0.84$
derived for the S0 NGC~4753
by Steiman-Cameron, Kormendy \& Durisen (1992)
on the basis of fitting an inclined,
precessing disk model to the complicated pattern of the
galaxy's dust lanes. Their remarkably good fit
is independent of $(c/a)_\rho$; the flattening
constraints are based on the assumption that
the gas is smoothly distributed and has completed at least 6
orbits at all radii.
Stable rings around galaxies are not observed to have random
orientations, but are found preferentially close to
the equatorial or polar planes, suggesting that the potential
may be oblate.
In particular, polar ring galaxies (PRGs) are surrounded
by rings of gas and stars in orbits nearly perpendicular to the
central stellar plane; these rings can extend to
20 disk scale lengths.
Since in an oblate potential closed ring orbits are elongated
along the polar axis and have speeds that vary with ring azimuth,
the shape and kinematics of a polar ring are excellent
extended probes of $(c/a)_\rho$.
Early kinematic analyses of three PRGs
produced axes ratios for the {\it potentials\/} of
$0.86 < (c/a)_\Phi < 1.05$ with uncertainties of 0.2
(Schweizer, Whitmore \& Rubin 1983, Whitmore, McElroy \& Schweizer 1987),
corresponding to $0.58 \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} (c/a)_\rho \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 1.15$
with very large uncertainties.
Subsequent studies using more detailed
mass models and higher quality data over a larger radial
range have narrowed the range for the dark mass to
$0.3 \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} (c/a)_\rho \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.6$ (Arnaboldi {\it et~al.}\ 1993, Sackett
{\it et~al.}\ 1994, Sackett \& Pogge 1995);
in each galaxy, $(c/a)_\rho$ is
similar to the inferred flattening of the central {\it stellar\/} body.
Measurements of $(c/a)_\rho$ for spiral galaxies are rarer,
more difficult, and sorely needed.
Assuming that gas disks evolve gravitationally toward a
discrete bending mode in tilted rigid halos,
Hofner and Sparke (1994) find that moderate halo flattening
of $0.6 \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} (c/a)_\rho \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.9$ can
reproduce the observed {H{\sc I}}\ warps of five spirals.
In principle, $(c/a)_\rho$~can also be constrained by
the flaring of the {H{\sc I}}\ layer; in the most detailed study of this type,
Olling and van~Gorkom (1995) obtain $0.2 < (c/a)_\rho < 0.8$
for the dark halo of the Sc NGC~4244.
Since non-gravitational energy sources may be responsible for
a substantial fraction of the vertical support of gas
(Malhotra 1995, and references therein), this measurement
may be an upper limit to $(c/a)_{\rho}$.
\section{Parting Caveats and a Puzzle}
Since the mass distribution in cluster galaxies
may be modified by the interactions and
violent relaxation that shape the evolving cluster potential,
we have restricted this review to relatively
isolated galaxies that are more likely to be dynamically relaxed.
Furthermore, we have largely ignored ellipticals,
the inner few kpc of which are thought to be responsible
for the strong lensing of distant QSOs and radio sources.
Although selection effects operate to favor
flattened lenses in multiply-imaged systems (Kassiola \& Kovner 1993),
image inversion techniques yield lenses
that are surprisingly flattened (Kochanek 1995a, and
references therein) --- the {\it projected\/} $(c/a)_{\Phi} \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.8$
corresponds to $(c/a)_\rho \mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 0.4$.
Can these flat lenses be reconciled with the axis ratio distribution of
ellipticals, which peaks at $q = 0.7$ (Ryden 1992), or are disk galaxies
implicated?
\vskip 0.3cm
\noindent\small{
P.D.S. gratefully acknowledges travel support from
the Leids Kerkhoven-Bosscha Fonds and the International Astronomical Union.
}
|
\section{Introduction}
It is well-known that symmetry plays an important role in modern physics.
Gauge symmetry leads to the Standard Model in high energy physics;
crystallographic space symmetry is fundamental to solid state physics,
conformal symmetry is crucial to string theory and critical phenomenon.
In a sense, the progress of modern physics is accompanied by study
of symmetry.
The mathematical counterpart of all the above motioned symmetries
and other popular ones in physics is group. Recently there has
been great interest in the study of quantum
group and quantum symmetry. Quantum group in contrast to its literal
meaning is not a group, even not a semi-group. However quantum group
is the deformation of the universal enveloping algebra of a
finite-dimensional semi-simple Lie algebra introduced by Drinfel'd
\cite{Drinfeld86} and Jimbo \cite{Jimbo85} in their study of the Yang-Baxter
equation (YBE). In 1967, Yang \cite{Yang}
discovered that if a certain consistency condition is satisfied, the quantum
mechanical many body problem on a line with the potential $\displaystyle
c\sum_{i<j}\delta(x_i-x_j)$
can be solved exactly by making use of Bethe Ansatz.
The consistency condition in Yang's paper and the commutivity
condition for the transfer matrix of the eight-vertex model in
statistical physics found by Baxter \cite{Baxter72a} are referred as
the YBE \cite{Faddeev79,Faddeev82}.
At this stage we would like to draw readers attention that the
word ``quantum'' in quantum group is from the YBE: solutions of
the classical YBE are closely related with classical or semi-simple group,
while solutions of the quantum YBE related with quantum group.
The ``quantum'' thus really differs from the canonical quantization and
possesses different meanings for different systems.
Conventionally, quantum group is the Hopf algebra which is neither commutative
nor cocommutative. A Hopf algebra is endowed with the algebra homomorphisms:
comultiplication $\Delta$ and counit $\epsilon$, and the algebra
anti-homomorphism: antipode $S$.
As long as $q$ is not a root of unity, the representation theory of
quantum group runs parallel to the classical theory \cite{Rosso88}.
The representations
are labeled by the same highest weight vectors. Most of the standard
expressions valid for classical algebras have $q$-analogs which often amount
to
replacing ordinary number by $q$-numbers. However, if $q$ is a root of
unity, representation theory changes drastically
\cite{Concini89}--\cite{Pasquier90}. This is one of
the most intriguing situations which are absent in the ``classical'' case.
The highest weight representations are still well defined, however, they are
no longer irreducible in general. Many of representations appearing in
the decomposition of tensor products of irreducible representations
are reducible but not fully reducible \cite{Pasquier90}. If $q$ is a root
of unity,
quantum group contains a large ``standard'' central subalgebra, and new
representations appear, the so-called cyclic representations
\cite{Concini89}.
Quantum group (quantum symmetry) is now a popular topic in different fields of
modern physics due to its richer structure than that of Lie group.
Quantum group is used as the natural structure to characterize and classify
rational conformal field theory \cite{Alvarez89}--\cite{Moore89}.
Integrable lattice models are constructed and solved by quantum group
theory. In Hamiltonian systems, quantum group is an enlarged symmetry
that maintains invariance of equations of motion. In quantum group
approach, a consistent description of vibrating and rotating molecular
spectra is given. Quantum group is also crucial in knot theory
\cite{Resh87a}--\cite{Lee92}, quantum optics \cite{Chai90}--\cite{CChang}
and gauge field theory \cite{Chau93}.
Quantum group theory has been developed in different directions. In the
quantum space approach \cite{Manin88,Manin87}, the initial object is a
quadratic algebra which is considered being
as the polynomial algebra on a quantum linear space. Quantum group appears
like a group of automorphisms of the quantum linear space.
The crucial ingredient of the matrix pseudogroup approach
\cite{Woronowicz87a}-- \cite{Woronowicz88}
states that any commutative $C^*$ algebra with unit element
is isomorphic to an algebra of all continuous functions on some compact
topological manifolds.
Other important approaches can be found in References
\cite{Reshitikhin89,Faddeev,Wess-Zumino}.
We do not discuss all of them and relations between
them, merely restrict ourselves to investigating quantum group as
quantized
universal enveloping algebra, because most of the applications of quantum
group in modern physics are expressed in this approach.
This presentation is by no means an exhaustive overview of this rapid
developing
subject. Likewise the references we give have no pretense to
completeness. Further materials can be found, for example, in References
\cite{Jimbo}--\cite{Ma93} and references therein.
This article is organized as follows. In Sec.$2$, we start to briefly
review the subject of the YBE, indeed the attempts of solving the YBE
motivated Drinfel'd and Jimbo to introduce the theory of quantum group.
Sec.$3$ begins with stating the basic facts of Hopf algebras,
the language in which the quantum group theory is written. After defining the
suitable form of quantization for Lie bi-algebra, quantum group
is presented as quantum double. In Sec.$4$, the representation theory of
quantum group is discussed in detail, for both generic and
non-generic $q$. Both the highest weight and cyclic representations are
investigated. Sec.$5$---Sec.$8$ are devoted to discussing
applications of quantum group, i.e., quantum symmetry in modern physics.
In Sec.$5$, by means of symplectic geometry, it is shown that quantum group
can be realized in a Hamiltonian system -- symmetric top system. The
Hamiltonian system, in which quantum group is realized, obeys the same
equation of motion as the standard Hamiltonian system, but these two
Hamiltonian systems have different constants of motion. In Sec.$6$, the
integrable lattice model are mapped into the XXZ spin chain through the
very anisotropic limit. Quantum symmetry in the model is shown. By
constructing a basis of a primitive left ideal from the Young operators
of quantum group, we transform the Schr\"{o}dinger equation of the model
to be a set of coupled linear equations which can be solved by the standard
method. In Sec.$7$, the quantum symmetry of the minimal model and the WZNW
model in their Coulomb gas version, is formulated in terms of a type of
screened vertex operators, which define the representation space of a
quantum group. The conformal properties of these operators exhibit a deep
interplay between the quantum group and the Virasoro algebra. In Sec.$8$
we give a complete treatment of vibrating and
rotating molecules in quantum group theoretic approach. The energy spectra of
the model possessing
quantum symmetry exhibit the properties of the infrared,
vibrational and rotational Raman spectra as well as vibrational and
rotational structures of electronic transitions of molecules. The exact
selection rules and wave functions are obtained. The Taylor expansion of
the analytic formulas of the approach reproduces
Dunham expansion.
\section{Yang-Baxter equation}
\subsection{Integrable quantum field theory}
For over two decades the YBE has been studied as
the master equation in lattice statistical physics and integrable quantum
field theory. The YBE first manifested itself in the work of Yang
\cite{Yang}. He considered a one dimensional quantum mechanical many body
problem with Hamiltonian
\begin{equation}
H=-\sum_{i=1}^N\frac{\partial^2}{\partial x_i^2}+2c\sum_{i<j}
\delta(x_i-x_j)~,~~~c>0~.
\end{equation}
Use is making of Bethe Ansatz, to write the wave function of the system
in the following form
\begin{equation}
\psi=\sum_PA_P(Q)\exp(ik_Px_Q)~,
\end{equation}
where $k_Px_Q\equiv\displaystyle\sum_{i=1}^Nk_{p_i}x_{q_i}$, $k_i$ is a
set of unequal numbers, $x_i$ is the position of the $i$-th particle and
satisfies
$0<x_{q_1}<x_{q_2}\cdots<x_{q_N}<L$,
$P=\left(p_1,p_2, \cdots ,p_N\right)$
and
$Q=\left(q_1,q_2, \cdots ,q_N \right)$ are two permutations of $N$ bodies.\\
Then we get the eigenenergy of the system as
\begin{equation}
E=\sum_{i=1}^Nk_i^2~.
\end{equation}
The solution of the problem thus reduces to obtain $k_i~(i=1,~2,~\cdots,~N)$
and $A_P(Q)$. Let us introduce the operator $T_R$ by
\begin{equation}
T_RA_P(Q)\equiv A_P(QR)~.
\end{equation}
For convenience, we denote $T_{ij}=T_{(ij)}, ~T_i=T_{(i~i+1)}$. \\
For the continuity of $\psi$,
\begin{equation}
A_{P}(Q)+A_{P'}(Q)=A_{P}(Q')+A_{P'}(Q')
\end{equation}
and the discontinuity of its derivative,
\begin{equation}
i(k_{p_{i+1}}-k_{p_i})\left(A_{P}(Q)
-A_{P'}(Q')\right)=c\left(A_{P}(Q)+A_{P'}(Q)\right)~,
\end{equation}
to be satisfied as required by the $\delta$-type interaction between
particles, it is sufficient to demand
\begin{equation}
\begin{array}{rcl}
A_{P}(Q)&=&\displaystyle\frac{i(k_{p_{i+1}}-k_{p_i}) T_i+c}
{i(k_{p_{i+1}}-k_{p_i})-c}A_{P'}(Q)\\[4mm]
&\equiv&Y_i(k_{p_{i+1}}-k_{p_i})A_{P'}(Q)~,
\end{array}
\end{equation}
where we have used the notation
$Y_i(u)\equiv\displaystyle\frac{iuT_i+c}{iu-c}$,
$P'$ and $Q'$ are two permutations of $N$ bodies and related with $P$ and $Q$
through $P'=P(i~i+1)$ and $Q'=Q(i~i+1)$. \\
Using the operator $Y_i(u)$ repeatedly, we can express $A_P(Q)$ in terms of
$A_E(Q)$, where $E$ is the unit element of the permutation group $S_N$.
For example, when $P=(321)$ we have
\begin{equation}
\begin{array}{rcl}
A_{(321)}(Q)&=&Y_1(k_2-k_3)Y_2(k_1-k_3)Y_1(k_1-k_2)A_{(123)}(Q)\\
&=&Y_2(k_1-k_2)Y_1(k_1-k_3)Y_2(k_2-k_3)A_{(123)}(Q)~.
\end{array}
\end{equation}
{}From the above equation, we see that $Y_1$ and $Y_2$ must satisfy the
following consistency condition
\begin{equation}
Y_1(k_2-k_3)Y_2(k_1-k_3)Y_1(k_1-k_2)=Y_2(k_1-k_2)Y_1(k_1-k_3)Y_2(k_2-k_3)~.
\end{equation}
In general case, the consistency condition possesses the form
\begin{equation}\label{0.yb1}
\begin{array}{l}
Y_i(u)Y_j(v)=Y_j(v)Y_i(u)~,~~~~|i-j| \geq 2~,\\
Y_i(u)Y_i(-u)=1~,\\
Y_i(u)Y_{i+1}(u+v)Y_i(v)=Y_{i+1}(v)Y_i(u+v)Y_{i+1}(u)~.
\end{array}
\end{equation}
Imposing the periodic boundary condition of $\psi$, \\ \\
$\displaystyle\sum_PA_P(q_1,q_2,\cdots,q_N)
\exp \left(i\left(k_{p_2}x_{q_2}+k_{p_3}x_{q_3}+\cdots
+k_{p_N}x_{q_N}\right)\right)$
\begin{equation}\label{0.07}
=\sum_PA_P(q_2,q_3,\cdots,q_N,q_1)\exp
\left(i\left(k_{p_1}x_{q_2}+k_{p_2}x_{q_3}+\cdots
+k_{p_{N-1}}x_{q_N}+k_{p_N}L\right)\right)~,
\end{equation}
we obtain\\ \\
$Y_1(k_1-k_i)Y_2(k_2-k_i)\cdots Y_{i-1}(k_{i-1}-k_i)A_E(Q)$
\begin{equation}
=\exp\left(ik_iL\right)T_1T_2\cdots T_{N-1}Y_{N-1}(k_i-k_N)Y_{N-2}
(k_i-k_{N-1})\cdots Y_{i}(k_i-k_{i+1})A_E(Q)~.
\end{equation}
Defining
\begin{equation}\label{0.yb3}
\begin{array}{l}
R_{i(i+1)}(u)\equiv T_iY_i(u)=\displaystyle\frac{u-icT_i}{u+ic}~,\\[4mm]
R_{ij}(u) \equiv \displaystyle\frac{u-icT_{ij}}{u+ic} ~,
\end{array}
\end{equation}
we have, similar with Eq.(\ref{0.yb1}),
\begin{equation}\label{0.YB4}
\begin{array}{l}
R_{ij}(u)R_{kl}(v)=R_{kl}(v)R_{ij}(u)~,\\
R_{ij}(u)R_{ij}(-u)=1~,
\end{array}
\end{equation}
and
\begin{equation}\label{0.yb5}
R_{ij}(u)R_{ik}(u+v)R_{jk}(v)=R_{jk}(v)R_{ik}(u+v)R_{ij}(u)~.
\end{equation}
Equation (\ref{0.yb5}) is just the YBE in integrable quantum field
theory. Yang provided first solution of the YBE, Eq.(\ref{0.yb3}).
In terms of $R_i$, the periodic boundary condition (\ref{0.07})
can be cast into the form
\begin{equation}
{\cal M}_iA_E(Q)=e^{ik_iL}A_E(Q),
\end{equation}
where
$${\cal M}_i\equiv R_{(i+1)i}(k_{i+1}-k_i)\cdots
R_{Ni}(k_{N}-k_i)R_{1i}(k_{1}-k_i)\cdots
R_{(i-1)i}(k_{i-1}-k_i)~.$$
It is straightforward to show that the operators ${\cal M}_i$ commute with
each other. As a result, $A_E(Q)$ is a common eigenstate
of the operators ${\cal M}_i$ with eigenvalue $e^{ik_iL}$. The problem
thus reduces to an eigenvalue problem for $k_i$. This eigenvalue problem
can be solved by a second use of Bethe Ansatz \cite{Yang}.
\subsection{Lattice statistical physics}
Independently, the YBE has arisen in Baxter's papers \cite{Baxter72a},
\cite{Baxter73a}--\cite{Baxter82} as the commutivity
condition for the transfer matrices of the so called eight-vertex model
in lattice statistical physics.
Consider a two-dimensional square lattice \cite{Baxter82}
(Fig.1). Degrees of freedom of
vertex model are associated with links of the lattice and
interact at the vertices (Fig.2).
For a horizontal row of the lattice with the adjacent vertex edges,
let $\alpha=\{\alpha_i,\cdots,\alpha_n\}$ be the state variables on the
lower row of the vertical edges, $\alpha'=\{\alpha'_i,\cdots,\alpha'_n\}$
be the
state variables on the upper row, and $\beta=\{\beta_i,\cdots,\beta_n\}$
be the state variables on the horizontal edges (Fig.3).
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(30,40)
\multiput(25,10)(0,5){5}{\line(1,0){25}}
\multiput(27.5,7.5)(5,0){5}{\line(0,1){25}}
\put(27.5,5.5){1}
\put(32.5,5.5){2}
\put(47.5,5.5){n}
\put(23,10){1}
\put(23,15){2}
\put(23,30){m}
\end{picture}
\hspace{10em}Fig.1. Square lattice.
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(20,30)
\put(30,10){\line(1,0){10}}
\put(35,5){\line(0,1){10}}
\put(28,10){i}
\put(35,3){j}
\put(41,10){k}
\put(35,16){l}
\end{picture}
\hspace{7em}Fig.2. Vertex $w(i,j,k,l)$, the Boltzmann
\hspace{9.5em} weight of the vertex model.
\thicklines
\setlength{\unitlength}{10pt}
\begin{picture}(50,15)(20,3)
\put(25,10){\line(1,0){30}}
\multiput(27.5,7)(5,0){6}{\line(0,1){6}}
\put(27.5,5){$\alpha_1$}
\put(32.5,5){$\alpha_2$}
\put(47.5,5){$\alpha_{n-1}$}
\put(52.5,5){$\alpha_{n}$}
\put(27.5,14){$\alpha'_1$}
\put(32.5,14){$\alpha'_2$}
\put(47.5,14){$\alpha'_{n-1}$}
\put(52.5,14){$\alpha'_{n}$}
\put(25.5,10.5){$\beta_1$}
\put(30.5,10.5){$\beta_2$}
\put(35.5,10.5){$\beta_3$}
\put(50.5,10.5){$\beta_{n}$}
\put(55.5,10.5){$\beta_{n+1}=\beta_1$}
\label{fig1}
\end{picture}
\hspace{4.5em}Fig.3. Row-to-Row transfer matrix ${\cal V}^{(n)}$ for the Vertex
model.
It is convenient to adopt a Hamiltonian picture, in which ``time'' flows
upward
on the lattice, and the various configurations of vertical links are considered
as independent possible states of the system at a given time. Time evolution
is carried out by the row-to-row transfer matrix ${\cal V}^{(n)}(u)$,
whose matrix elements ${\cal V}^{(n)}_{\alpha,\alpha'}(u)$ is defined by
\begin{equation}\label{0.2.11}
{\cal V}^{(n)}_{\alpha,\alpha'}(u)=\sum_{\beta_1\cdots\beta_n}
w(\beta_1,\alpha_1,\beta_2,\alpha_1'|u)
w(\beta_2,\alpha_2,\beta_3,\alpha_2'|u)
\cdots w(\beta_n,\alpha_n,\beta_1,\alpha_n'|u)~,
\end{equation}
where $w(\beta_i,\alpha_i,\beta_i',\alpha_i'|u)$ is Boltzmann weight
of the vertex.
In terms of the transfer matrix ${\cal V}^{(n)}(u)$, the partition function
$Z_N$ and the free energy per site $f$ are given by
\begin{equation}
\begin{array}{l}
Z_N=\displaystyle\sum_{\alpha^{(1)}\alpha^{(2)}\cdots \alpha^{(m)}}
{\cal V}^{(n)}_{\alpha^{(1)},\alpha^{(2)}}(u) {\cal V}^{(n)}_{\alpha^{(2)},
\alpha^{(3)}}(u)\cdots {\cal V}^{(n)}_{\alpha^{(m-1)},\alpha^{(m)}}(u)
{\cal V}^{(n)}_{\alpha^{(m)},\alpha^{(1)}}(u)={\rm Tr}\left({\cal
V}^{(n)}(u)\right)^m~,\\
f=-k_BT\displaystyle\lim_{N\to\infty}N^{-1}\log Z_N~,
\end{array}
\end{equation}
where $N=mn$ is the number of lattice sites and summation is taken over all
configurations of arrows. \\
Let ${\cal V}^{(n)}(u+v)$ be another transfer matrix where the Boltzmann
weight $w(\beta_i,\alpha_i,\beta_{i+1},\alpha_i'|u)$ is
replaced by $w(\beta_i,\alpha_i,\beta_{i+1},\alpha_i'|u+v)$. From
Eq.(\ref{0.2.11}),
we have
\begin{equation}\label{0.2.13} \begin{array}{rcl}
\left({\cal V}^{(n)}(u){\cal V}^{(n)}(u+v)\right)_{\alpha,\alpha'}&=&
\displaystyle\sum_{\gamma}{\cal V}^{(n)}(u)_{\alpha,\gamma}
{\cal V}^{(n)}(u+v)_{\gamma,\alpha'}\\
&=&\displaystyle\sum_{\beta\bar{\beta}}\prod_{i=1}^{n}
X(\beta_i,\bar{\beta}_i|\beta_{i+1}, \bar{\beta}_{i+1}|\alpha_i,\alpha'_i)~,
\end{array}
\end{equation}
where
\begin{equation}\label{0.2.14}
X(\beta_i,\bar{\beta}_i|\beta_{i+1}, \bar{\beta}_{i+1}|\alpha_i,\alpha_i')
=\sum_{\gamma_i} w(\beta_i,\alpha_i,\beta_{i+1},\gamma_i|u)
w(\bar{\beta}_i,\gamma_i,
\bar{\beta}_{i+1},\alpha_i'|u+v)~.
\end{equation}
Equation (\ref{0.2.13}) can be written as a compact form
\begin{equation}\label{0.2.15}
\left({\cal V}^{(n)}(u){\cal V}^{(n)}(u+v)\right)_{\alpha,\alpha'}={\rm
Tr}X(\alpha_1,\alpha'_1)X(\alpha_2,\alpha'_2)
\cdots X(\alpha_n,\alpha'_n)~,
\end{equation}
where $X(\alpha,\alpha')$ is the matrix with element
$X(\beta_i,\bar{\beta}_i|\beta_{i+1},\bar{\beta}_{i+1}|\alpha_i,\alpha_i')$
in row $(\beta_i,\bar{\beta}_i)$ and in column
$(\beta_{i+1},\bar{\beta}_{i+1})$.
Similarly, we define $X'$ with
$w(\beta_i,\alpha_i,\beta_{i+1},\alpha_i'|u)$
and $w(\beta_i,\alpha_i,\beta_{i+1},\alpha_i'|u+v)$ interchanged in
Eq.(\ref{0.2.14}), which lead to
\begin{equation}\label{0.2.16}
({\cal V}^{(n)}(u){\cal V}^{(n)}(u+v))_{\alpha,\alpha'}={\rm
Tr}X'(\alpha_1,\alpha_1')X'(\alpha_2,\alpha_2')
\cdots X'(\alpha_n,\alpha_n')~.
\end{equation}
{}From Eqs.(\ref{0.2.15}) and (\ref{0.2.16}), we see that ${\cal V}^{(n)}(u)$
and ${\cal V}^{(n)}(u+v)$ commute if there exists a matrix $W$ satisfying
\begin{equation}\label{0.0010}
X(\alpha_i,\alpha_i')=WX'(\alpha_i,\alpha_i')W^{-1}~.
\end{equation}
The matrix $W$ has rows labeled by $(\beta_i,\bar{\beta}_i)$, and columns
labeled by $(\beta_{i+1},\bar{\beta}_{i+1})$ with elements
$w(\bar{\beta}_i,\beta_i,\beta_{i+1},\bar{\beta}_{i+1}|v)$.
Equation (\ref{0.0010}) can then be cast into the following
form\\ \\ $\displaystyle\sum_{\gamma_i\beta_i''\bar{\beta}_i''}
w(\beta_i,\alpha_i,\beta_i'',\gamma_i|u)
w(\bar{\beta}_i,\gamma_i,\bar{\beta}_i'',\alpha_i'|u+v)
w(\bar{\beta}''_i,\beta''_i,\beta'_i,\bar{\beta}_i'|v)$
\begin{equation}\label{0.2.18}
=\sum_{\gamma_i\beta''_i\bar{\beta}''_i}
w(\bar{\beta}_i,\beta_i,\beta''_i,\bar{\beta}''_i|v)
w(\beta''_i,\alpha_i,\beta'_i,\gamma_i|u+v)w(\bar{\beta}''_i,\gamma_i,
\bar{\beta}'_i,\alpha'_i|u)~.
\end{equation}
This is the YBE in lattice statistical physics.\\
\subsection{Yang-Baxter equation}
When we identify
\begin{equation}
\begin{array}{l}
w(\beta_i,\alpha_i,\beta''_i,\gamma_i|u)=R^{\beta_i\beta''_i}_{\alpha_i\gamma_i}
(u)~,\\
w(\bar{\beta}_i,\gamma_i,\bar{\beta}''_i,\alpha'_i|u+v)=
R^{\bar{\beta}_i\bar{\beta}''_i}_{\gamma_i\alpha'_i}(u+v)~,\\
w(\bar{\beta}''_i,\beta''_i,\beta'_i,\bar{\beta}'_i|v)=
R^{\bar{\beta}''_i\beta'_i}_{\beta''_i\bar{\beta}'_i}(v)~,
\end{array}
\end{equation}
Eq.(\ref{0.2.18}) becomes Eq.(\ref{0.yb5}). Furthermore, the quantities
$R$ and $w(\beta,\alpha,\beta',\alpha')$ can be obviously
interpreted as an operator ${\cal R}$ in the tensor product space
$V^{\otimes 2}$.
In the space $V^{\otimes 3}$ we introduce three
operators ${\cal R}_{12}$, ${\cal R}_{13}$, ${\cal R}_{23}$
corresponding to the three canonical embeddings of $V^{\otimes 2}$ into
$V^{\otimes 3}$ (for example, ${\cal R}_{12}={\cal R}\otimes I$,
${\cal R}_{23}=I\otimes {\cal R}$). Then Eqs.(\ref{0.yb5}), (\ref{0.2.18})
can be rewritten as
\begin{equation}\label{0.cd1}
{\cal R}_{12}(u){\cal R}_{13}(u+v){\cal R}_{23}(v)={\cal R}_{23}(v)
{\cal R}_{13}(u+v){\cal R}_{12}(u)~.
\end{equation}
The variables $u$ is called as the spectral parameter, and a solution of
Eq.(\ref{0.cd1}) is called as $R$ matrix.\\
The solution of the YBE corresponding to integrable quantum field theory
\cite{Yang} is
\begin{equation}\label{0.ys}
\begin{array}{rcl}
{\cal R}(u,c)&=&\displaystyle\frac{u-icP}{u+ic}\\[4mm]
&=&\displaystyle\frac{1}{u+ic}\left[
\begin{array}{cccc}
u-ic& 0 & 0 & 0 \\
0 & u &-ic& 0 \\
0 &-ic& u & 0 \\
0 & 0 & 0 &u-ic \end{array}\right]~,
\end{array}
\end{equation}
where $P$ is the transposition operator in $V\otimes V$,
\begin{equation}
\begin{array}{rcl}
P:~~V_1\otimes V_2&\to& V_2\otimes V_1~,\\
P:~~a\otimes b&\to& b\otimes a~.
\end{array}
\end{equation}
In 6-vertex model, the solution \cite{Baxter82} has the form
\begin{equation}\label{0.ts}
{\cal R}(u)=\rho\left[\begin{array}{cccc}
\sin(\eta+u)& & & \\
& \sin u &\sin\eta& \\
& \sin\eta&\sin u & \\
& & &\sin(\eta+u)
\end{array}\right]~,
\end{equation}
where
$$\begin{array}{l}
\rho\sin(\eta+u)=\exp(-i\beta\epsilon_1)~,\\
\rho\sin u=\exp(-i\beta\epsilon_3)~,\\
\rho\sin\eta=\exp(-i\beta\epsilon_5)~,
\end{array}$$
$\epsilon_i$ ($i=1,3,5$) are three distinct energies of 6-vertex model, and
$\eta$ is a free parameter.\\
Another type of solution \cite{Baxter73a} is
\begin{equation}
{\cal R}(u)=\left[\begin{array}{cccc}
a(u)& & &d(u)\\
&b(u)&c(u)&\\
&c(u)&b(u)&\\
d(u)& & &a(u)\end{array}\right]~,
\end{equation}
where
$$
\begin{array}{rcl}
a(u)&=&\theta_0(\eta)\theta_0(u)\theta_1(\eta+u)~,\\
b(u)&=&\theta_0(\eta)\theta_1(u)\theta_0(\eta+u)~,\\
a(u)&=&\theta_1(\eta)\theta_0(u)\theta_0(\eta+u)~,\\
a(u)&=&\theta_1(\eta)\theta_1(u)\theta_1(\eta+u)~,
\end{array}$$
and $\theta_i(u)$ are the elliptic theta functions
$$\begin{array}{rcl}
\theta_0(u)&=&\displaystyle\prod_{i=1}^{\infty}\left(1-2p^{i-1/2}\cos 2\pi u
+p^{2i-1}\right)(1-p^i)~,\\
\theta_1(u)&=&2p^{1/8}\sin\pi u\displaystyle\prod_{i=1}^{\infty}
\left(1-2p^{i}\cos 2\pi u
+p^{2i}\right)(1-p^i)~,
\end{array}$$
$\eta$ and $p$ are free parameters.\\
Solutions of the YBE on high genus Riemann surface ($g>1$) have been found
recently \cite{Perk87,Perk88}.\\
Making use of the basis for $V$
\begin{equation}
\begin{array}{l}
{\cal R}(u)=\sum{\cal R}_{ij}^{kl}E_{ik}\otimes E_{jl}~,\\
E_{ij}=(\delta_{im}\delta_{jn})_{m,n=1,2,\cdots,{\cal D}}~,
\end{array}
\end{equation}
we see that the YBE (\ref{0.cd1}) amounts to ${\cal D}^6$ homogeneous nonlinear
equations with ${\cal D}^4$ unknowns ${\cal R}_{ij}^{kl}$. Here we have
used the notation ${\cal D}\equiv {\rm dim}V$. Generally speaking, it is very
difficult to solve the YBE \cite{Jimbo89,Vega90} save for certain limiting
cases.
\subsection{Classical Yang-Baxter equation}
There is a set of important solutions (quasi-classical solutions) of the
YBE which contains an
extra parameter $\gamma$ (quantization parameter). We expand these
solutions near zero point of the quantization parameter $\gamma$
\begin{equation}
{\cal R}(u,\gamma)=({\rm scalar})\times\left(I+\gamma r(u)+O(\gamma^2)\right)~.
\end{equation}
The $r(u)$ is called as the classical limit of ${\cal R}(u,\gamma)$.
In the zero $\gamma$ limit, the YBE (\ref{0.cd1}) reduces to the
classical Yang-Baxter equation (CYBE) for $r(u)$
\cite{Kulish82,Drinfeld88,Tyan83},
\begin{equation}\label{0.cd2}
[r_{12}(u),r_{13}(u+v)]+[r_{12}(u),r_{23}(v)]+[r_{13}(u+v),r_{23}(v)]=0~.
\end{equation}
The CYBE that is in Lie bracket form has one less parameter than the YBE.
It is easier to discuss solutions of the CYBE.\\
Let $g$ be a Lie algebra,
and $r(u)$ be a $g\otimes g$-valued function. In terms of a basis
$\{X_\mu\}$ of $g$, we write
\begin{equation}
r(u)=\sum_{\mu,\nu}r^{\mu\nu}(u)X_\mu\otimes X_\nu
\end{equation}
with $c$-valued functions $r^{\mu\nu}(u)$. Let further
\begin{equation}
r_{12}(u)=\sum_{\mu,\nu}r^{\mu\nu}X_{\mu}\otimes X_\nu\otimes I~~~~\in
\left(U(g)\right)^{\otimes 3}~,
\end{equation}
and so on, where $U(g)$ denotes the universal enveloping algebra of $g$.
Then each term in Eq.(\ref{0.cd2}) actually is in $g^{\otimes
3}$, for example,
\begin{equation}
[r_{12}(u),r_{23}(v)]=\sum r^{\mu\nu}(u)r^{\rho\sigma}(v)X_\mu\otimes[X_\nu,X_
\rho]\otimes X_\sigma~.
\end{equation}
For arbitrary triplet of representations $(\pi,V_i)~(i=1,~2,~3)$ of
$g$, $(\pi_i\otimes\pi_j)~(r_{ij}(u))$ yields a matrix solution of the CYBE.
The following are two familiar examples of solutions of the CYBE.\\
\begin{itemize}
\item Rational Solution. \\
For an orthonormal basis $\{X_\mu\}$ of $g$ with a non-degenerate invariant
bilinear form $\Omega$, we have a solution of the CYBE
\begin{equation}\label{0.rs}
\begin{array}{l}
r(u)=\displaystyle\frac{1}{u}\Omega~,\\
\Omega\equiv\displaystyle\sum_{\mu}X_{\mu}\otimes X_{\mu}~,
\end{array}
\end{equation}
namely the rational solution. This solution corresponds
to that given by Eq.(\ref{0.ys}). The $c=0$ limit of Eq.(\ref{0.ys}) is
\begin{equation}
r(u)=\frac{1}{u}P=\frac{1}{2u}\left(1+\sum_{i=1}^3\sigma_i\otimes
\sigma_i\right)~,
\end{equation}
where $\sigma_i$ are Pauli matrices. As a results
\begin{equation} \label{0.su2}
r(u)=\frac{1}{u}\sum_{i=1}^3\left(\frac{\sigma_i}{2}\right)\otimes
\left(\frac{\sigma_i}{2}\right)~,
\end{equation}
is a solution of the CYBE. It is well-known that
$\displaystyle\frac{\sigma_i}{2}$ are
generators of the fundamental representation of the Lie algebra $su(2)$.
Equation (\ref{0.rs})
is a generalization of Eq.(\ref{0.su2}) to general Lie algebra. \\ \\ \\
\item Trigonometric Solution.\\
In Cartan-Weyl basis, the non-degenerate invariant bilinear form $\Omega$ has
the form
\begin{equation}
\begin{array}{rcl}
\Omega&=&\displaystyle\sum_j H_j\otimes H_j+\displaystyle\sum_{\alpha\in
\Delta_+}\left(E_{\alpha}\otimes E_{-\alpha}+E_{-\alpha}\otimes
E_{\alpha}\right)\\
&\equiv&C_0+C_++C_-~,
\end{array}
\end{equation}
where $\Delta_+$ is the positive root space of $g$,
$$\begin{array}{l}
C_0=\displaystyle\sum_jH_j\otimes H_j~,\\
C_+=\displaystyle\sum_{\alpha\in\Delta_+}E_\alpha\otimes E_{-\alpha}~,\\
C_-=\displaystyle\sum_{\alpha\in\Delta_+}E_{-\alpha}\otimes E_{\alpha}~.
\end{array}$$
One can readily verify that
\begin{equation}
r=C_0+2C_+
\end{equation}
is a trigonometric solution of the CYBE.
\end{itemize}
The solutions of the CYBE can be classified completely by the following
theorem \cite{Belavin82}.\\
{\em The non-degenerate solution $r(u)$ extends
meromorphically to the whole plane
${\bf C}$, with all simple poles; $r(u)$ is classified by
$\Gamma=\{ {\rm the~set~of ~poles~of~r(u)} \}$:\\
\begin{itemize}
\item Elliptic function has rank($\Gamma)=2$ and exists only for $g=sl(n)$.
\item Trigonometric function has rank($\Gamma)=1$ and exists for each type,
and can be classified by making use of the Dynkin diagram for affine Lie
algebra.
\item Rational function has rank($\Gamma)=0$.
\end{itemize} }
Provided a solution $r(u)\in
g\otimes g$ of the CYBE, one may ask whether exists a quasi-classical
${\cal R}(u,\gamma)$ having $r(u)$ as its classical limit. It is this
``quantization'' problem that has motivated Drinfel'd and Jimbo to introduce
the theory of quantum group.
\section{Quantum group theory}
\subsection{Hopf algebra}
An algebra over the field ${\bf C}$ is a linear space $A$ which possesses two
linear operations: the multiplication $m$ and the unit $\eta$
\begin{equation}
\begin{array}{rcl}
& & m:~~~A\otimes A\to A~,\\
& & m(a\otimes b)=ab~,~~~~a,b\in A~,\\
& & \eta:~~~{\bf C}\to A~,
\end{array}
\end{equation}
and satisfies the following axioms
\begin{equation}
\begin{array}{rcl}
& & m(m\otimes{\rm id})=m({\rm id}\otimes m)~,\\
& & m({\rm id}\otimes\eta)=m(\eta\otimes {\rm id})={\rm id}~,
\end{array}
\end{equation}
i.e., it is associative (Fig.4) and has unit element (Fig.5). When we deal with
tensor
product representations of the underlying abstract algebra, another linear
operation, namely comultiplication $\Delta$, should be endowed.
The comultiplication satisfies the following relations
\begin{equation}
\begin{array}{rcl}
& &\Delta(ab)=\Delta(a)\Delta(b)~,\\
& &(\Delta\otimes{\rm id})\Delta=({\rm id}\otimes\Delta)\Delta~,
\end{array}
\end{equation}
i.e., $\Delta$ is an algebra homomorphism and co-associative (Fig.6). After
introducing the comultiplication, we can compose two representations ( for
example, add angular momenta). This takes place when two physical systems, each
of them is within a certain representation, interact. \\
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,26)(0,2)
\put(25,20){\vector(3,1){15}}
\put(20,17){$A\otimes A\otimes A$}
\put(41,24){$A\otimes A$}
\put(41,10){$A\otimes A$}
\put(48,25){\vector(3,-1){15}}
\put(25,16){\vector(3,-1){15}}
\put(48,11){\vector(3,1){15}}
\put(62,17){$A$}
\put(28,22.3){$m$}
\put(30,23){$\otimes$}
\put(32,23.67){id}
\put(28,12){id}
\put(30,11.3){$\otimes$}
\put(32,10.7){$m$}
\put(56,23){$m$}
\put(56,11.5){$m$}
\end{picture}
\hspace{13.5em}Fig.4. Associativity.
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,30)(0,8)
\put(25,15){\vector(1,0){15}}
\put(63,15){\vector(-1,0){15}}
\put(44,17){\vector(0,1){15}}
\put(18,14){${\bf C}\otimes A$}
\put(41,14){$A\otimes A$}
\put(64,14){$A\otimes {\bf C}$}
\put(43.2,32.5){$A$}
\put(21,17){\line(6,5){19}}
\put(22,17){\line(6,5){19}}
\put(67,17){\line(-6,5){19}}
\put(68,17){\line(-6,5){19}}
\put(28.5,15.7){$\eta\otimes$id}
\put(51.5,15.7){id$\otimes\eta$}
\put(44.7,23){$m$}
\end{picture}
\hspace{14.5em}Fig.5. The unit.
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,35)(0,5)
\put(40,25){\vector(-3,-1){15}}
\put(20,17){$A\otimes A\otimes A$}
\put(41,24){$A\otimes A$}
\put(41,10){$A\otimes A$}
\put(63,20){\vector(-3,1){15}}
\put(40,11){\vector(-3,1){15}}
\put(63,16){\vector(-3,-1){15}}
\put(62,17){$A$}
\put(28,22.3){$\Delta$}
\put(30,23){$\otimes$}
\put(32,23.67){id}
\put(28,12){id}
\put(30,11.3){$\otimes$}
\put(32,10.7){$\Delta$}
\put(56,23){$\Delta$}
\put(56,11.5){$\Delta$}
\end{picture}
\hspace{13em}Fig.6. Co-associativity.
It is well-known that the operation $\eta$ signals the existence of a
unit in $A$. In the same manner we now define an operation $\epsilon$,
called the counit (Fig.7), $\epsilon: ~A\to {\bf C}$, satisfying
\begin{equation}\label{1.1.4}
({\rm id}\otimes \epsilon)\Delta=(\epsilon\otimes{\rm id})\Delta={\rm id}~.
\end{equation}
If all the above axioms be satisfied, the set $(A,m,\Delta,\eta,\epsilon)$ is
called a bi-algebra. Being related with the fact that every element of a group
has an inverse, certain bi-algebras possess an extra operation, antipode $S$
(Fig.8), $S:~A\to A$, with the properties
\begin{equation}\label{1.1.5}
m(S\otimes{\rm id})\Delta=m({\rm id}\otimes S)\Delta=\eta\circ\epsilon~.
\end{equation}
A bi-algebra with antipode is called a Hopf algebra \cite{Able}. We now
illustrate a useful example of Hopf algebras. \\
Let $g$ denote a Lie algebra and $U(g)$ denote its universal enveloping
algebra; if we define
\begin{itemize}
\item the multiplication $m$ as the ordinary multiplication in $U(g)$,
\item $\Delta(x)=x\otimes 1+1\otimes x~,~~~~\forall x\in g~,$
\item $\eta(\alpha)=\alpha {\bf 1}~,$
\item $\epsilon({\bf 1})=1$ and zero on all other elements,
\item $S(x)=-x~,$
\end{itemize}
$U(g)$ becomes a Hopf algebra.
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,33)(2,10)
\put(40,15){\vector(-1,0){15}}
\put(48,15){\vector(1,0){15}}
\put(44,32){\vector(0,-1){15}}
\put(18,14){${\bf C}\otimes A$}
\put(41,14){$A\otimes A$}
\put(64,14){$A\otimes {\bf C}$}
\put(43.2,32.5){$A$}
\put(21,17){\line(6,5){19}}
\put(22,17){\line(6,5){19}}
\put(67,17){\line(-6,5){19}}
\put(68,17){\line(-6,5){19}}
\put(28.5,15.7){$\epsilon\otimes$id}
\put(51.5,15.7){id$\otimes\epsilon$}
\put(44.7,23){$\Delta$}
\label{fig4}
\end{picture}
\hspace{13em}Fig.7. The counit.
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,46)(-5,5)
\put(25,20){\vector(3,1){15}}
\put(22,17){$A\otimes A$}
\put(41,24){$A\otimes A$}
\put(41,10){$A\otimes A$}
\put(48,25){\vector(3,-1){15}}
\put(25,16){\vector(3,-1){15}}
\put(48,11){\vector(3,1){15}}
\put(63.5,17){$A$}
\put(28,22.3){$S$}
\put(30,23){$\otimes$}
\put(32,23.67){id}
\put(28,12){id}
\put(30,11.3){$\otimes$}
\put(32,10.7){$S$}
\put(56,23){$m$}
\put(56,11.5){$m$}
\put(6,18){\vector(1,0){15}}
\put(3,17){$A$}
\put(32.5,33){${\bf C}$}
\put(4,19){\line(0,1){15}}
\put(65,34){\vector(0,-1){14}}
\put(4,34){\vector(1,0){28}}
\put(35,34){\line(1,0){30}}
\put(12.5,18.7){$\Delta$}
\put(18,34.7){$\epsilon$}
\put(49,34.7){$\eta$}
\end{picture}
\hspace{12em}Fig.8. The antipode.
Strictly speaking, the above definitions for
$\Delta,~\eta,~\epsilon$ and $S$ are merely restricted on the subset
$g$ of $U(g)$. Note that $g$
is not a Hopf algebra because it is not an associative algebra. However, it
is readily to see that these operations can be extended in a unique
manner to all of $U(g)$
so that the Hopf algebra axioms are satisfied everywhere. The universal
enveloping algebras are commutative Hopf algebras, i.e., we have the equality
$\Delta=P\circ\Delta$.
If a Hopf algebra $(A,m,\Delta,\eta,\epsilon,S)$ is not cocommutative, i.e.,
$\Delta'\equiv P\circ\Delta\not=\Delta$, then it is not difficult to verify
that $(A,m,\Delta',\eta,\epsilon,S')$ is also a Hopf algebra (so called the
related Hopf algebra of $(A,m,\Delta,\eta,\epsilon,S)$), where we have used the
notation $S'=S^{-1}$.
\subsection{Quantization of Lie bi-algebra}
A Lie algebra $g$ with a co-antisymmetric operation
$\psi:~g\to g\otimes g~~~~ (P\circ\psi=-\psi$) satisfying the so
called 1-cocycle condition
$$\begin{array}{rcl}
& &[X,Y\otimes Z]\equiv [X,Y]\otimes Z+Y\otimes [X,Z]~,~~~~X,Y,Z\in g~,\\
& &\psi\left([X,Y]\right)=[\psi(X),Y]+[X,\psi(Y)]~,
\end{array}$$
is a Lie bi-algebra. A quantization of a Lie bi-algebra
$(A_{(0)},m_{(0)},\Delta_{(0)},\eta_{(0)},\epsilon_{(0)},\psi$)
\cite{Drinfeld86,Jimbo85,Jimbo86b,Manin88,Woronowicz89} is a noncommutative
algebra
$(A,m,\Delta,\eta,\epsilon)$ over the ring ${\bf C}[[\gamma]]$ ($\gamma$ is
the
quantization parameter). The space $A$ is the set of polynomials in $\gamma$
with coefficients in $A_0$. To factorize $\gamma A$ out is then
equivalent to
set $\gamma=0$, which corresponds to the classical limit. A new
noncommutative multiplication $m$ (denoted as $m(a\otimes b)=a\star b$) is of
the form
\begin{equation}
a\star b=\sum_{i=0}^{\infty}f_i(a,b)\gamma^i~,
\end{equation}
where $f_i:~A\otimes A\to A$. \\
Denote the canonical quotient map $A\to A_0$ by $\pi$, a complete
description of the
quantization $(A,m,\Delta,\eta,\epsilon)$ of $(A_{(0)},m_{(0)},\Delta_{(0)},
\eta_{(0)}, \epsilon_{(0)},\psi)$ is
\begin{itemize}
\item $A/\gamma A\cong A_{(0)}~,$
\item $(\pi\otimes \pi)\circ\Delta=\Delta_{(0)}\circ\pi~,$
\item $m_{(0)}\circ(\pi\otimes\pi)=\pi\circ m~,$
\item $\pi\circ\eta=\eta_{(0)}~,$
\item $\epsilon\circ\pi=\epsilon_{(0)}~,$
\item $\psi(\pi(a))=\pi\displaystyle\left(\frac{1}{\gamma}\left(\Delta(a)-
P\circ\Delta(a)\right)\right)~,~~~~\forall a\in A~.$
\end{itemize}
In the following, we will consider the details of the quantization of the
universal enveloping algebra $U(su(2))$.
There are two subalgebras in $U(su(2))$, i.e., the positive Borel
subalgebra,
$U^+(su(2))$, and the negative Borel subalgebra, $U^-(su(2))$,
generated by $X^+,~H$ and $X^-,~H$, respectively.
We first discuss the quantization of $U^+(su(2))$. Similar discussions can be
given to $U^-(su(2))$.
It is not difficult to verify that the map $\psi$ defined by
\begin{equation}
\psi(H)=0~,~~~ ~~~\psi(X^+)=2(H\otimes X^+-X^+\otimes H)~,
\end{equation}
and the ordinary Lie bracket
\begin{equation}
[X^+,H]=X^+~,
\end{equation}
form a Lie bi-algebra. To quantize this Lie bi-algebra, the first step is
to find out
the comultiplication on the set of polynomials in $\gamma$. In the classical
limit the comultiplication $\Delta$ must reduce to the ordinary
comultiplication
$\Delta_{(0)}$ on $U^+(su(2))$ and satisfy
\begin{equation}\label{1.ti1}
\psi(\pi(a))=\pi\left(\frac{1}{\gamma}\left(\Delta(a)-P\circ\Delta
(a)\right)\right)~.
\end{equation}
Also, it must be co-associative. The general form of the comultiplication is
\begin{equation}
\Delta=\sum_{i=0}^{\infty}\frac{\gamma^i}{i!}\Delta_{(i)}~.
\end{equation}
{}From its classical limit, we obtain
\begin{equation}
\begin{array}{rcl}
& & \Delta_0(H)=H\otimes 1+1\otimes H~,\\
& & \Delta_0(X^+)=X^+\otimes 1+1\otimes X^+~.
\end{array}
\end{equation}
The explicit form of Eq.(\ref{1.ti1}) reads
\begin{equation}\label{1.ti2}
\begin{array}{rcl}
& &\psi(H)=0=\Delta_1(H)-P\circ\Delta_{(1)}(H)~,\\
& &\psi(X^+)=2(H\otimes X^+-X^+\otimes H)=\Delta_{(1)}(X^+)-
P\circ\Delta_{(1)}(X^+)~.
\end{array}
\end{equation}
An obvious nontrivial solution of these equations is given by
\begin{equation}\label{1.ti3}
\begin{array}{rcl}
& &\Delta_{(1)}(H)=0~,\\
& &\Delta_{(1)}(X^+)=(H\otimes X^+-X^+\otimes H)~.
\end{array}
\end{equation}
The co-associativity of the comultiplication gives us a recursion relation
\begin{equation}\label{1.ti4}
\sum_{k=0}^{i}\left(\begin{array}{c}
i\\
k\end{array}\right)\left(\Delta_{(k)}\otimes 1-1\otimes
\Delta_{(k)}\right)\Delta_{(i-k)}=0~,
\end{equation}
where we have used the notation $\displaystyle\left(\begin{array}{c}
n\\
m\end{array}\right)
=\frac{n!}{m!(n-m)!}~.$\\
{}From Eqs.(\ref{1.ti1}) $\sim$ (\ref{1.ti4}), we get
\begin{equation}
\begin{array}{rcl}
& &\Delta_{(i)}(H)=0~,~~~(i\geq 1)~,\\
& &\Delta_{(i)}(X^+)=H^i\otimes X^++(-1)^iX^+\otimes H^i~.
\end{array}
\end{equation}
Making use of the above results, we obtain the comultiplication $\Delta$
\begin{equation}
\begin{array}{rcl}
& &\Delta(H)=H\otimes 1+1\otimes H~,\\
& &\Delta(X^+)=X^+\otimes q^{-H}+q^{H}\otimes X^+~,
\end{array}
\end{equation}
where the new quantization parameter $q\equiv e^\gamma$ is introduced. \\
Use is made of axioms of Hopf algebra,
to determine the form of other operations on the quantized universal
enveloping algebra $U_q^+\left(su(2)\right)$. One of these axioms is
\begin{equation}
\Delta(a\star b)=\Delta(a)\star\Delta(b)~,~~~~\forall a,b\in A~.
\end{equation}
In our context this means that the equality
\begin{equation}
\Delta\left([H,X^+]\right)=\left[\Delta(H),\Delta(X^+)\right]~,
\end{equation}
must be hold. It is easy to verify that
\begin{equation}
\left[\Delta(H),\Delta(X^+)\right]=\Delta(X^+)~,
\end{equation}
so that the ordinary $su(2)$ relation $[H,X^+]=X^+$ still holds after
quantization. \\ Equation (\ref{1.1.4}) reads
\begin{equation}
\begin{array}{rcl}
& &(\epsilon\otimes{\rm id})\Delta(H)=\epsilon(H)1+H=H~,\\
& &(\epsilon\otimes{\rm id})\Delta(X^+)=\epsilon(X^+)q^{-H}+
\epsilon(q^{H})X^+=X^+~.
\end{array}
\end{equation}
One of the solutions of these equations is
\begin{equation}
\epsilon(H)=0~,~~~~~~\epsilon(X^+)=0~.
\end{equation}
In the same manner the equations,
\begin{equation}
S(H)=-H~,~~~~~~S(X^+)=-q^{-1}X^+~,
\end{equation}
follow from the explicit form of Eq.(\ref{1.1.5})
\begin{equation}
\begin{array}{rcl}
& &m(S\otimes{\rm id})\Delta(H)=S(H)+H=0~,\\
& &m(S\otimes{\rm id})\Delta(X^+)=S(X^+)q^{-H}+S(q^H)X^+=0~.
\end{array}
\end{equation}
Defining
\begin{equation}\label{1.225}
\begin{array}{l}
t=q^{2H}~,~~~~tt^{-1}=1=t^{-1}t~,\\
e=q^HX^+~,
\end{array}
\end{equation}
we rewrite the Hopf algebra $U^+_q(su(2))$ as
\begin{equation}
\begin{array}{rcl}
& &tt^{-1}=1=t^{-1}t~,\\
& &tet^{-1}=q^2e~,\\
& &\Delta(e)=e\otimes 1+t\otimes e~,~~~~\Delta(t)=t\otimes t~,\\
& &\epsilon(e)=0~,~~~~\epsilon(t)=1~,\\
& &S(e)=-t^{-1}e~,~~~~S(t)=t^{-1}~.
\end{array}
\end{equation}
Similarly, in terms of a new set of definitions
\begin{equation}
\begin{array}{rcl}
& &\bar{t}=q^{2H}~,~~~~\bar{t}\bar{t}^{-1}=1=\bar{t}^{-1}\bar{t}~,\\
& &f=X^-q^{-H}~,
\end{array}
\end{equation}
the Hopf algebra $U_q^-(su(2))$ possesses the form
\begin{equation}
\begin{array}{rcl}
& &{\bar t}{\bar t}^{-1}=1={\bar t}^{-1}{\bar t}~,\\
& &{\bar t}f{\bar t}^{-1}=q^{-2}f~,\\
& &\Delta(f)=f\otimes {\bar t}^{-1}+1\otimes f~,~~~~
\Delta({\bar t})={\bar t}\otimes {\bar t}~,\\
& &\epsilon(f)=0~,~~~~\epsilon({\bar t})=1~,\\
& &S(f)=-f{\bar t}~,~~~~S({\bar t})={\bar t}^{-1}~.
\end{array}
\end{equation}
\subsection{Quantum double}
For a Hopf algebra $(A,~m,~\Delta,~\eta,~\epsilon,~S)$ the iterated
comultiplication $\Delta^{(n-1)}:~A\to A^{\otimes n}$ is inductively
defined by $\Delta^{(1)}=\Delta,~\Delta^{(n)}=(\Delta\otimes{\rm
id})\circ\Delta^{(n-1)}$.
Provided two Hopf algebras $A,~B$ and a non-degenerate bilinear form
\begin{equation}
\langle\cdots,\cdots\rangle:~A\times B\longrightarrow {\bf C}~,
\end{equation}
satisfying the following conditions
\begin{equation}\label{1.jc1}
\begin{array}{rcl}
& &\langle a^i,b_jb_k\rangle=\langle\Delta_A(a^i),b_j\otimes
b_k\rangle~,\\[2mm]
& &\langle a^ia^j,b_k\rangle=\langle a^j\otimes a^i,\Delta_B(b_k)
\rangle ~,\\[2mm]
& &\langle 1^A,b_i\rangle=\epsilon_B(b_i)~,~~~~~
\langle a^i,1_B\rangle=\epsilon_A(a^i)~,\\[2mm]
& &\langle S_A(a^i),S_B(b_j)\rangle=\langle a^i,b_j\rangle~,
\end{array}
\end{equation}
where $\Delta_{A~(B)},~\epsilon_{A~(B)},~S_{A~(B)}$ are the structure
operations of the Hopf algebra $A~(B),~1^{A~(B)}$ is its unit,
we have the following theorem.\\
{\em There exists a unique Hopf algebra $D$ with the following properties:
\begin{itemize}
\item $D$ contains $A,~B$ as Hopf subalgebras, i.e., $A$ and $B$ are
subalgebras of $D$, and
$$\begin{array}{rcl}
& &\Delta_D(a^i)=\Delta_A(a^i)~,~~~\epsilon_D(a^i)=\epsilon_A(a^i)~,~~~
S_D(a^i)=S_A(a^i)~,~~~~~
\forall a^i\in A~,\\
& &\Delta_D(b_i)=\Delta_B(b_i)~,~~~\epsilon_D(b_i)=\epsilon_B(b_i)~,~~~
S_D(b_i)=S_B(b_i)~,~~~~~
\forall b_i\in B~.
\end{array}$$
\item If $\{a^\alpha\}$, $\{b_\beta\}$ are bases of $A$ and $B$, the
product $\{a^\alpha b_\beta\}$ is a basis of $D$.
\item For $a^i\in A$ and $b_j\in B$
$$b_ja^i=\sum\langle a^{i(1)},S(b_{j(1)})\rangle \langle
a^{i(3)},b_{j(3)}\rangle
a^{i(2)}b_{j(2)}~~~\in D~,$$
where
$$\begin{array}{rcl}
& &\sum a^{i(1)}\otimes a^{i(2)} \otimes a^{i(3)} \equiv
\Delta_A^{(2)}(a^i)~,\\
& &\sum b_{j(1)}\otimes b_{j(2)} \otimes b_{j(3)} \equiv
\Delta_B^{(2)}(b_j)~.
\end{array}$$
The Hopf algebra $D$ is called as the quantum double of
$(A,~B,~\langle,\rangle)$ or simply, quantum double of $A$. \end{itemize}}
\subsubsection{$SU_q(2)$ as the quantum double}
For a nonzero complex number $q\in
{\bf C}^\times~({\bf C}\setminus\{0\}),~q^2\not=1$, consider the
two Hopf algebras $U_q^+\left(su(2)\right)$ and $U_q^-\left(su(2)\right)$.\\
Using the properties (\ref{1.jc1}) of the bilinear form
$\langle\cdots,\cdots\rangle$
repeatedly, we have
\begin{equation}
\begin{array}{rcl}
\langle te,\bar{t}\rangle&=&\langle e\otimes t,\bar{t}\otimes\bar{t}\rangle\\
&=&\langle e,\bar{t}\rangle \langle t,\bar{t}\rangle
{}~,\\
\langle te,\bar{t}\rangle&=&q^2\langle et,\bar{t}\rangle
=q^2\langle t\otimes e,{\bar t}\otimes{\bar t}\rangle
\\
&=&q^2\langle e,\bar{t}\rangle\langle t,{\bar
t}\rangle~.
\end{array}
\end{equation}
The equation
\begin{equation}
\begin{array}{rcl}
1&=&\langle t,1\rangle=\langle t,{\bar t}\cdot{\bar t}^{-1}\rangle\\
&=&\langle t,\bar{t}\rangle\langle t,\bar{t}^{-1}\rangle~,
\end{array}
\end{equation}
leads to
\begin{equation}
\langle t,{\bar t}\rangle=\langle t,{\bar t}^{-1}\rangle^{-1}\not=0~.
\end{equation}
Since $q^2\not= 1$, we must have
\begin{equation}
<e,{\bar t}>=0~.
\end{equation}
Similar calculations yield
\begin{equation}
<t,f>=0~.
\end{equation}
Making comparison of
\begin{equation}
\begin{array}{rcl}
\langle te,f\rangle&=&\langle e\otimes t,\Delta(f)\rangle
=\langle e,f\rangle\langle t,\bar{t}^{-1}\rangle+
\langle e,1\rangle\langle t,f\rangle \\
&=&\langle e,f\rangle \langle t,\bar{t}^{-1}\rangle ~,\\
\end{array}
\end{equation}
and
\begin{equation}
\begin{array}{rcl}
\langle te,f\rangle&=&q^2\langle et,f\rangle=q^2\langle t\otimes e,
\Delta(f)\rangle
=q^2\langle t,f\rangle\langle e,\bar{t}^{-1}\rangle+
q^2\langle t,1\rangle\langle e,f\rangle \\
&=&q^2\langle e,f\rangle ~,\\
\end{array}
\end{equation}
where $\langle e,f\rangle\not= 0$ for non-degeneracy, we obtain
\begin{equation}
\langle t,{\bar t}\rangle =q^{-2}=\langle t,\bar{t}^{-1}\rangle^{-1}~.
\end{equation}
In general,
\begin{equation}
\begin{array}{rcl}
\langle e^mt^n,f^{m'}{\bar t}^{n'}\rangle&=&\langle e^m,f^{m'}\rangle\langle
t,{\bar t}\rangle^{nn'}\\[2mm]
&=&\delta_{mm'}q^{m(m-1)/2}[m]!\langle e,f\rangle^m q^{-2nn'}~,
\end{array}
\end{equation}
where $[m]=\displaystyle\frac{q^m-q^{-m}}{q-q^{-1}}$, is a $q$-number,
$[m]!=[m][m-1]\cdots[2][1]$ is used. \\
The inner-product $\langle e,f\rangle$ can always be re-scaled by
the transformation $e \to ce$ ($c\in {\bf C}^\times$). For latter
convenience we choose
\begin{equation}\label{1.j3.3}
\langle e,f\rangle=\frac{-1}{q-q^{-1}}~.
\end{equation}
Now we are ready to examine the quantum double $D$ of
$(U_q^+\left(su(2)\right),~U_q^-\left( su(2)\right),~\langle,
\rangle)$. From the above theorem an element of $D$ is a unique linear
combination of monomials
$e^mt^nf^{m'}{\bar t}^{n'}~(m,~m'\in Z_{\geq 0},~
n,~n'\in Z)$ and the commutators of $e$ with $t$ or $f$ with ${\bar t}$
can be computed easily.
We now compute $ft,~\bar{t}e~,$and $fe$. Since
\begin{equation}
\begin{array}{rcl}
& &\Delta^{(2)}(t)=t\otimes t\otimes t~,\\
& &\Delta^{(2)}(\bar{t})=\bar{t}\otimes \bar{t}\otimes\bar{t}~,
\end{array}
\end{equation}
and
\begin{equation}
\begin{array}{rcl}
& &\Delta^{(2)}(e)=e\otimes 1\otimes 1+t\otimes e\otimes 1+t\otimes t\otimes
e~,\\
& &\Delta^{(2)}(f)=f\otimes{\bar t}^{-1}\otimes {\bar t}^{-1}+1\otimes f\otimes
{\bar t}^{-1}+1\otimes 1\otimes f~,
\end{array}
\end{equation}
we get
\begin{equation}
\begin{array}{rcl}
ft&=&\langle t,S(f)\rangle\langle t,{\bar t}^{-1}\rangle t{\bar t}^{-1}\\
& & +\langle t,S(1)\rangle\langle t,\bar{t}^{-1}\rangle tf
+\langle t,S(1)\rangle\langle t,f\rangle t\cdot 1\\
&=&q^2 tf~,\\[3mm]
\bar{t}e&=&\langle e,S(\bar{t})\rangle\langle 1,{\bar t}\rangle 1\cdot t\\
& & +\langle t,S(\bar{t})\rangle\langle 1,\bar{t}\rangle e\bar{t}
+\langle t,S(\bar{t})\rangle\langle e,\bar{t}\rangle t\bar{t}\\
&=&q^2 e\bar{t}~,
\end{array}
\end{equation}
and
\begin{equation}
\begin{array}{rcl}
fe&=&\langle e,S(f)\rangle\langle 1,{\bar t}^{-1}\rangle 1\cdot \bar{t}^{-1}\\
& & +\langle t,S(1)\rangle\langle 1,\bar{t}^{-1}\rangle ef
+\langle t,S(1)\rangle\langle e,f\rangle t\cdot 1\\
&=&-\langle e,f\rangle{\bar t}^{-1}+ef+\langle e,f\rangle t~.
\end{array}
\end{equation}
Using Eq.(\ref{1.j3.3}) we have
\begin{equation}
[e,f]=\frac{t-{\bar t}^{-1}}{q-q^{-1}}~.
\end{equation}
Setting $t=\bar{t}$ the quantum group $SU_q(2)$ follows
\begin{equation}\label{1.a1}
\begin{array}{l}
tet^{-1}=q^2e~,~~~~tft^{-1}=q^{-2}f~,\\[2mm]
[e,f]=\displaystyle\frac{t-t^{-1}}{q-q^{-1}}~,\\[2mm]
\Delta(e)=e\otimes 1+t\otimes e~,\\
\Delta(f)=f\otimes t^{-1}+1\otimes f~,\\
\Delta(t)=t\otimes t~,\\
\epsilon(e)=0=\epsilon(f)~,~~~~\epsilon(t)=1~,\\
S(e)=-t^{-1}e~,~~~~S(f)=-ft~,~~~~S(t)=t^{-1}~.
\end{array}
\end{equation}
In the literature there are available various versions of the quantum group.
Setting
\begin{equation}\label{1.a2}
e'=t^{-n}e~,~~~~f'=ft^n~,~~~~t'=t~,
\end{equation}
would result in
\begin{equation}\label{1.a3}
\begin{array}{l}
t'e'{t'}^{-1}=q^2e'~,~~~~t'f'{t'}^{-1}=q^{-2}f'~,\\[2mm]
[e',f']=\displaystyle\frac{t'-{t'}^{-1}}{q-q^{-1}}~,\\[2mm]
\Delta(e')=e'\otimes {t'}^{-n}+{t'}^{1-n}\otimes e'~,\\
\Delta(f')=f'\otimes {t'}^{n-1}+{t'}^n\otimes f~,\\
\Delta(t')=t'\otimes t'~,\\
\epsilon(e')=0=\epsilon(f')~,~~~~\epsilon(t')=1~,\\
S(e')=-t'^{-1}e'~,~~~~S(f')=-f't'~,~~~~S(t')=t'^{-1}~.
\end{array}
\end{equation}
It is easy to see from comparison between Eq.(\ref{1.a3}) and
Eq.(\ref{1.225}) that
$n=\displaystyle\frac{1}{2}$ case of Eq.(\ref{1.a3}) is the version in
Chevalley basis,
\begin{equation}\label{1.a4}
\begin{array}{l}
[H,X^\pm]=X^\pm~,\\[1mm]
[X^+,X^-]=[2H]~,\\[1mm]
\Delta(X^\pm)=X^\pm\otimes q^{-H}+q^H\otimes X^\pm~,\\
\Delta(H)=H\otimes 1+1\otimes H~,\\
\epsilon(X^\pm)=0=\epsilon(H)~,\\
S(X^\pm)=-q^{\mp 1}X^\pm~,~~~~S(H)=-H~.
\end{array}
\end{equation}
\subsubsection{$U_q(g)$ as the quantum double}
Now we come to discuss the general quantum group $U_q(g)$ as the quantum
double. Let $g$ denote an ordinary simple Lie algebra or untwisted affine
Kac-Moody algebra. The
corresponding generalized Cartan matrix $A=(a_{ij})_{1\leq i,j\leq l}$
($a_{ij}=\displaystyle\frac{1}{d_i}\langle\alpha_i,\alpha_j\rangle,~d_i=
\frac{1}{2}\langle\alpha_i,
\alpha_i\rangle)$ is symmetrizable in a sense that there exist nonzero $d_i$
satisfying $d_ia_{ij}=d_ja_{ji}$. For a nonzero complex number
$q$ ($q^{2d_i}\not=1$), in Chevalley basis, define
$U(g)$ as the associative ${\bf C}$-algebra with unity, with $3l$
generators
$$X_i^+~,~~~X_i^-~,~~~H_i~,~~~~~~(1\leq i\leq l)~.$$
Consider the standard Borel subalgebras of $U(g)$,\\
$U^+(g):~{\rm generated~by}~X^+_i~{\rm and}~H_i~~~(1\leq i\leq l)$,
\begin{equation}\label{1.6.15}
\begin{array}{l}
[H_i,H_j]=0~,\\[1mm]
[H_i,X_j^+]=a_{ij}X_i^+~,\\[1mm]
[X_i^+,X_j^+]=0~,~~~~~{\rm if}~a_{ij}=0~,\\[2mm]
\displaystyle\sum_{m=0}^{1-a_{ij}}(-1)^m\left(\begin{array}{c}
1-a_{ij}\\
m
\end{array}
\right)(X_i^+)^{1-a_{ij}-m}X_j^+(X_i^+)^m=0~,~~~~ (i\not=j)~,\\[2mm]
\Delta(H_i)=H_i\otimes 1+1\otimes H_i~,\\
\Delta(X_i^+)=X_i^+\otimes 1+1\otimes X_i^+~,\\
\epsilon(H_i)=0=\epsilon(X_i^+)~,\\
S(H_i)=-H_i~,~~~~S(X_i^+)=-X_i^+~;
\end{array}
\end{equation}
$U^-(g):~{\rm generated~by}~X^-_i~{\rm and}~H_i~~~~(1\leq i\leq l)$
\begin{equation}\label{1.6.151}
\begin{array}{l}
[H_i,H_j]=0~,\\[1mm]
[H_i,X_j^-]=-a_{ij}X_i^-~,\\[1mm]
[X_i^-,X_j^-]=0~,~~~~~{\rm if}~a_{ij}=0~,\\[2mm]
\displaystyle\sum_{m=0}^{1-a_{ij}}(-1)^m\left(\begin{array}{c}
1-a_{ij}\\
m
\end{array}
\right)(X_i^-)^{1-a_{ij}-m}X_j^-(X_i^-)^m=0~,~~~~ (i\not=j)~,\\[2mm]
\Delta(H_i)=H_i\otimes 1+1\otimes H_i~,\\
\Delta(X_i^-)=X_i^-\otimes 1+1\otimes X_i^-~,\\
\epsilon(H_i)=0=\epsilon(X_i^-)~,\\
S(H_i)=-H_i~,~~~~S(X_i^-)=-X_i^-~.
\end{array}
\end{equation}
The quantization of $U^+(g)$ and $U^-(g)$ yields the
multiplication $m$, comultiplication $\Delta$, counit $\epsilon$ and antipode
$S$ for $U^+_q(g)$
and $U^-_q(g)$ as,
\begin{equation}
\begin{array}{l}
\Delta(H_i)=H_i\otimes 1+1\otimes H_i~,\\
\Delta(X_i^+)=X_i^+\otimes q_i^{-H_i}+q_i^{H_i}\otimes X_i^+~,\\
\epsilon(X_i^+)=0=\epsilon(H_i)~, \\
S(X_i^+)=-q_i^{-1}X_i^+~,~~~~S(H_i)=-H_i~,
\end{array}
\end{equation}
and
\begin{equation}
\begin{array}{l}
\Delta(H_i)=H_i\otimes 1+1\otimes H_i~,\\
\Delta(X_i^-)=X_i^-\otimes q_i^{-H_i}+q_i^{H_i}\otimes X_i^-~,\\
\epsilon(X_i^-)=0=\epsilon(H_i)~, \\
S(X_i^-)=-q_iX_i^-~,~~~~S(H_i)=-H_i~,
\end{array}
\end{equation}
where $q_i=q^{d_i}$.\\
It is convenient to introduce the operators $e_i,~t_i,~$ and $f_i,~
\bar{t}_i~~~~(1\leq i\leq l)$ as
\begin{equation}
\begin{array}{l}
e_i=q_i^{H_i}X_i^+~,~~~~t_i=q_i^{2H_i}~,\\
f_i=X_i^-q_i^{-H_i}~,~~~~\bar{t}_i=q_i^{2H_i}~.
\end{array}
\end{equation}
To avoid confusion the new operators $x_i$ and $y_i$ are introduced by
$$\begin{array}{l}
x_i=e_i~,~~~~(1\leq i\leq l)~,\\
y_i=f_i~,~~~~(1\leq i\leq l)~.\end{array}$$
For $\beta\in I$ ($I=\displaystyle\bigoplus_{i=1}^lZ_{\geq
0}\alpha_i$) define
$$U_\beta^+={\rm
linear~span~of~}\{x_{i_1},~\cdots,~x_{i_r}|\alpha_{i_1}
+\cdots\alpha_{i_r}=\beta\}~,~~~(\beta\in I)~,$$
$$U_{-\beta}^-={\rm linear~span~of~}\{y_{i_1},~\cdots,~y_{i_r}|\alpha_{i_1}
+\cdots\alpha_{i_r}=\beta\}~,~~~(\beta\in I)~.$$
Then, $U_{\pm \beta}^\pm$ is finite dimension for each $\beta\in I$.
Similar to the case for $SU_q(2)$, there is a general theorem for the
quantum double $D$ of $U_q^+(q)$.
{\em
\begin{itemize}
\item There exists a unique non-degenerate bilinear pairing $\langle\cdots,
\cdots\rangle:~ U_q^+(g)\times U_q^-(g)\to {\bf C}$ so that
$$\begin{array}{l}
\langle t_i,\bar{t}_j\rangle=q_i^{-a_{ij}}~,~~~\langle
t_i,f_i\rangle=0~,\\[2mm]
\langle e_i,\bar{t}_i\rangle=0~,~~~\langle e_i,f_i\rangle=-
\delta_{ij}\displaystyle\frac{1}{q_i-q_i^{-1}}~.
\end{array}$$
\item The restriction of the pairing $\langle\cdots,\cdots\rangle$ to
$U_{\beta}^+\times
U_{-\beta}^-$ is non-degenerate for each $\beta\in I$.
\end{itemize} }
It can be shown that $U_q(g)$ follows from the quantum double
$D$ of $U_q^+(g)$ by setting $t_i={\bar t}_i$
\begin{equation}
D/\langle t_i-{\bar t}_i\rangle\simeq U_q(g)~.
\end{equation}
The resultant quantum group $U_q(g)$ has the form \cite{Drinfeld85}--
\cite{Lusztig90c}
\begin{equation}\label{1.6.152}
\begin{array}{rcl}
& &t_it_j=t_jt_i~,~~~~~t_it_i^{-1}=t_i^{-1}t_i=1~,\\
& &t_ie_jt_i^{-1}=q_i^{a_{ij}}e_j~,~~~~t_if_jt_i^{-1}=q_i^{-a_{ij}}f_j~,\\[2mm]
& &[e_i,f_j]=\delta_{ij}\displaystyle\frac{t_i-t_i^{-1}}
{q_i-q_i^{-1}}~,\\[3mm]
& &\displaystyle\sum_{m=0}^{1-a_{ij}}(-1)^m\left[\begin{array}{c}
1-a_{ij}\\
m\end{array}
\right]_{q_i}(e_i)^{1-a_{ij}-m}e_j(e_i)^m=0~,~~~~
(i\not=j)~,\\
& &\displaystyle\sum_{m=0}^{1-a_{ij}}(-1)^m\left[\begin{array}{c}
1-a_{ij}\\
m\end{array}
\right]_{q_i}(f_i)^{1-a_{ij}-m}f_j(f_i)^m=0~,~~~~
(i\not=j)~,\\
& &\Delta(t_i)=t_i\otimes t_i~,\\
& &\Delta(e_i)=e_i\otimes 1+t_i\otimes e_i~,~~~~
\Delta(f_i)=f_i\otimes t_i^{-1}+1\otimes f_i~,\\
& &\epsilon(e_i)=0=\epsilon(f_i)~,~~~~\epsilon(t_i)=1~,\\
& &S(e_i)=-t_i^{-1}e_i~,~~~~S(f_i)=-f_it_i^{-1}~,~~~~S(t_i)=t_i^{-1}~,
\end{array}
\end{equation}
where we have used the notation $\displaystyle\left[\begin{array}{c}
m\\
n\end{array}\right]_q=\frac{[m]_q!}{[n]_q![n-m]_q!}$. \\
In Chevalley basis, the quantum group $U_q(g)$ can be written as
\begin{equation}
\begin{array}{l}
[H_i,H_j]=0~,\\[1mm]
[H_i,X_j^\pm]=\pm a_{ij}X_i^\pm~,\\[1mm]
[X_i^+,X_j^-]=\delta_{ij}[H_i]_{q_i}~,\\[1mm]
[X_i^\pm,X_j^\pm]=0~,~~~~~{\rm if}~a_{ij}=0~,\\[2mm]
\displaystyle\sum_{m=0}^{1-a_{ij}}(-1)^m\left[\begin{array}{c}
1-a_{ij}\\
m
\end{array}
\right]_{q_i}(X_i^\pm)^{1-a_{ij}-m}X_j^\pm(X_i^\pm)^m=0~,~~~~
(i\not=j)~,\\[2mm]
\Delta(H_i)=H_i\otimes 1+1\otimes H_i~,\\
\Delta(X_i^\pm)=X_i^\pm\otimes q_i^{H_i}+q_i^{-H_i}\otimes X_i^+~,\\
\epsilon(H_i)=0=\epsilon(X_i^\pm)~,\\
S(H_i)=-H_i~,~~~~S(X_i^\pm)=-q^{-\rho} X_i^\pm q^{\rho}~,
\end{array}
\end{equation}
where $\rho=\displaystyle\sum_i H_i$.
\subsubsection{Universal $R$-matrix}
For the sake of definiteness, let us assume that $A,~B$ are finite-dimensional
Hopf algebras, and $\{a^i\}$ and $\{b_i\}$ are the bases of the
Hopf algebras $A$ and $B$ respectively. With respect to the non-degenerate
bilinear form $\langle\cdots,\cdots\rangle$,
$$\langle a^i,b_j\rangle=\delta_{ij}~.$$
Suppose that there exists another set of similar bases of the Hopf
algebras $A$ and $B$
$\{{a'}^{i}\}$ and $\{b'_{i}\}$
$${a'}^{i}=\sum_k a^kM_{ki}~,~~~~b'_{i}=\sum_lN_{il}b_l~,$$
\begin{equation}
\delta_{ij}=\langle {a'}^{i},b'_{j}\rangle=\sum_{kl}M_{ki}N_{jl}\langle a^k,
b_l\rangle=\sum_kN_{jk}M_{ki}~,
\end{equation}
i.e., $N$ is the inverse of $M$, $N=M^{-1}$.\\
Then we get
\begin{equation}
\sum_i{a'}^{i}\otimes b'_{i}=\sum_{ikl}M_{ki}N_{il}a^k\otimes b_l
=\sum_ia^i\otimes b_i~,
\end{equation}
and the element,
\begin{equation}
{\cal R}=\sum_i a^i\otimes b_i~~~\in A\otimes B\subset D\otimes D~,
\end{equation}
is {\it independent} of choice of the dual bases. The element ${\cal
R}$ is the so-called universal $R$ matrix. There is a theorem for the
universal $R$ matrix.
{\em
\begin{itemize}
\item ${\cal R}\Delta(x)=\Delta'(x){\cal R}~,~~~~\forall x\in D~,~~~~\Delta'
=P\circ\Delta~$,
\item $(\Delta\otimes{\rm id}){\cal R}={\cal R}_{13}{\cal R}_{23}~,~~~~
({\rm id}\otimes\Delta){\cal R}={\cal R}_{13}{\cal R}_{12}~.$
\item $(\epsilon\otimes{\rm id}){\cal R}=1=({\rm id}\otimes \epsilon){\cal R}~,
{}~~~~(S\otimes{\rm id}){\cal R}={\cal R}^{-1}=({\rm id}\otimes S^{-1}){\cal
R}~,$
\end{itemize}}
where
$$\begin{array}{rcl}
& &{\cal R}_{12}=\displaystyle\sum_i a^i\otimes b_i\otimes 1~,\\
& &{\cal R}_{13}=\displaystyle\sum_i a^i\otimes 1\otimes b_i~,\\
& &{\cal R}_{23}=\displaystyle\sum_i 1\otimes a^i\otimes b_i~.
\end{array}$$
In particular, ${\cal R}$ is invertible in $D\otimes D$. The pair $(D,{\cal
R})$
is the so-called quasi-triangular Hopf algebra. There is a
relationship of the quasi-triangular Hopf algebra to the YBE. \\
Making Comparison of
\begin{equation}
\begin{array}{rcl}
({\rm id}\otimes\Delta'){\cal R}&=&{\cal R}_i^{(1)}\otimes\Delta'({\cal
R}_i^{(2)})
={\cal R}_i^{(1)}\otimes{\cal R} \Delta({\cal R}
_i^{(2)}){\cal R}^{-1}\\
&=&({\rm id}\otimes{\cal R})\left({\cal R}_i^{(1)}
\otimes\Delta
({\cal R}_i^{(2)})\right)({\rm id}\otimes{\cal
R}^{-1})
={\cal R}_{23}({\rm id}\otimes\Delta){\cal R}{\cal
R}
_{23}^{-1}\\
&=&{\cal R}_{23}{\cal R}_{13}{\cal R}_{12}
{\cal R}_{23}^{-1}
\end{array}
\end{equation}
with
\begin{equation}
\begin{array}{rcl}
({\rm id}\otimes\Delta'){\cal R}&=&({\rm id}\otimes P)({\rm
id}\otimes\Delta){\cal R}
=({\rm id}\otimes P){\cal R}_{13}{\cal R}_{12}\\
&=&{\cal R}_{12}{\cal R}_{13}~,
\end{array}
\end{equation}
yields
\begin{equation}
{\cal R}_{12}{\cal R}_{13}{\cal R}_{23} ={\cal R}_{23} {\cal R}_{13} {\cal
R}_{12}~.
\end{equation}
This is precisely the YBE. \\
Now we are discussing the explicit form of the universal
$R$ matrix for $SU_q(2)$. Notice that the Hopf algebras
$A=U_q^+\left(su(2)\right)$ and
$B=U_q^-\left(su(2)\right)$ are not finite dimension. However if we suppose
that
$q$ is a primitive $p$-th root of unity for an odd integer $p>1$, it is not
difficult to verify that
$$e^p=0=f^p~,~~~~t^p=1=\bar{t}^p~,$$
so that the quotient algebras
\begin{equation}
\bar{A}=A/\langle e^p,t^p-1\rangle~,~~~~\bar{B}=B/\langle
f^p,\bar{t}^p-1\rangle~,
\end{equation}
are finite dimension, and
$$a^{mn}=e^mt^n~,~~~~b_{mn}=f^m\bar{t}^n~,~~~~(0\leq m,n\leq p)~,$$
are linear bases.\\
Since
\begin{equation}
\langle a^{mn},b_{m'n'}\rangle=\delta_{mm'}q^{m(m-1)/2}[m]!\left(\frac{-1}
{q-q^{-1}}\right)^mq^{-2nn'}~,
\end{equation}
$\{b_{mn}\}$ is not a dual basis of $\{a^{mn}\}$. The basis dual to
$\{a^{mn}\}$ should be given by
\begin{equation}
a_{mn}^*=\frac{q^{-m(m-1)/2}}{[m]!}\left(-(q-q^{-1})\right)^m\frac{1}{p}
\sum_{n'=0}^{p-1}q^{2nn'}b_{mn'}~.
\end{equation}
As a result, the explicit form of the universal $R$ matrix is
\begin{equation}
\begin{array}{rcl}
& &{\cal R}=\bar{\cal C}{\cal K}~,\\
& &\bar{\cal C}=\displaystyle\sum_{m=0}^{p-1}
\frac{q^{-m(m-1)/2}}{[m]!}\left(-(q-q^{-1})\right)^m
e^m\otimes f^m~,\\
& &{\cal K}=\displaystyle\frac{1}{p}\sum_{n,n'=0}^{p-1}q^{2nn'}t^n
\otimes\bar{t}^{n'}~.
\end{array}
\end{equation}
For the pair $(\pi,V),~(\pi',V')$ of representations of $D$, if
$$t\cdot v=q^\mu v~,~~~~~\bar{t}\cdot v'=q^\nu v'~,~~~~~(v\in V,~v'
\in V',~\mu,\nu\in Z) ~,$$
we have
\begin{equation}
\begin{array}{rcl}
{\cal K}\cdot v\otimes v'&=& q^{-\mu\nu/2}v\otimes v'\\
&=&q^{-2H\otimes H}v\otimes v'~.
\end{array}
\end{equation}
Therefore we can write the universal $R$ matrix as
\begin{equation}
{\cal R}=\sum_{m=0}^{p-1}\frac{q^{-m(m-1)/2}}{[m]!}\left(-(q-q^{-1})\right)^m
(e^m\otimes f^m)\cdot q^{-2H\otimes H}~.
\end{equation}
In terms of the $q$-exponential
$$\exp_qz=\sum_{m=0}^\infty\frac{q^{-m(m-1)/2}}{[m]!}z^m~,$$
with the condition that $e^p=0=f^p$,
the universal $R$ matrix can be formally written as
\begin{equation}
{\cal R}=\exp_q\left(-(q-q^{-1})e\otimes f\right)q^{-2H\otimes H}~.
\end{equation}
This expression for universal $R$ matrix is independent of $p$, and in
fact it is also valid for generic $q$.
The explicit form of the universal $R$ matrix except for the case
$SU_q(2)$ is more complicated \cite{Reshetikhin91}--\cite{Drinfeld90}.
\section{Representation theory}
As the same as the representation theory of group, a linear representation
of a quantum group $A$ is a pair $(\pi,V)$ consisting of a vector space $V$
and a homomorphism
$$\pi:~~A\to {\rm End}(V)~,~~~~~\pi(ab)=\pi(a)\pi(b)~,$$
where, as usual, End$(V)$ denotes the space of all linear maps from $V$ to
itself.
We often drop $\pi$ and write $\pi(a)v$ as $av$ ($a\in A,~v\in V)$. A
representation $(\pi,V)$ of a quantum group $A$ is called irreducible if $V$
has not non-trivial submodules, i.e., if $W(\subset V)$ is a subspace such
that
$A\cdot W\subset W$, either $W=\{0\}$, or $W=V$. If the vector space $V$
is finite dimension, the representation $(\pi,V)$ is a finite-dimensional
representation. For two representations $(\pi,V),~(\pi',V')$, an intertwiner
is the map $\phi:~V\to V'$, which commutes with the action of $A$, namely
$\phi\circ\pi(a)=\pi'(a)\circ\phi$ for all $a\in A$ (Fig.9). We say
that $(\pi,V)$ and $(\pi',V')$ are equivalent if there exists an intertwiner
that is an isomorphism. In other words, if there exist such bases of
$V,~V'$ that the matrices representing $\pi(a)$ and $\pi'(a)$ are
identical for all $a\in A$.
In this section we discuss the inequivalent irreducible finite-dimensional
representations of quantum group. When dealing with representations , we
only use the algebra structure, ignoring the Hopf algebra structure
$(\Delta,\epsilon,S)$.
The Hopf algebra structure enters into the picture when the relationship
between various representations is to be involved.
\setlength{\unitlength}{5pt}
\begin{picture}(50,25)
\put(30,18){\vector(1,0){15}}
\put(27,17){$V$}
\put(46,17){$V'$}
\put(36.5,19){$\phi$}
\put(28,16){\vector(0,-1){10}}
\put(47,16){\vector(0,-1){10}}
\put(30,4){\vector(1,0){15}}
\put(27,3){$V$}
\put(46,3){$V'$}
\put(36.5,5){$\phi$}
\put(22,10){$\pi(a)$}
\put(48,10){$\pi'(a)$}
\end{picture}
\hspace{7em}Fig.9. Intertwiner of representations.
\subsection{Representations for generic $q$}
\subsubsection{Representations of $SU_q(2)$}
It is well-know that the quantum group $SU_q(2)$ is generated by the
generators: $e,~f$ and $t^\pm$, and is subject to
\begin{equation}\label{2.cha1}
\begin{array}{rcl}
& &tt^{-1}=1=t^{-1}t~,\\[2mm]
& &tet^{-1}=q^{2}e~,~~~~tft^{-1}=q^{-2}f~,\\[2mm]
& &[e,f]=\displaystyle\frac{t-t^{-1}}{q-q^{-1}}~.
\end{array}
\end{equation}
Defining $v$ be the non-zero eigenvector of $e$ in the representation
$(\pi,V)$
$$ev=\lambda v,~~~~~v\in V,~\lambda\in {\bf C}~,$$
and making use of Eq.(\ref{2.cha1}), we obtain
$$e(t^{\pm 1}v)=q^{\mp 2}\lambda(t^{\pm 1}v)~.$$
Since $t$ is invertible, $t^{\pm 1} v\not=0$, and $q^{\mp 2}\lambda$
is also an eigenvalue of $e$ with the eigenstate $t^{\pm 1} v$.
If $\lambda\not=0$, we would have infinitely many eigenvalues and
eigenstates
$(\lambda,v),~(q^{\mp 2}\lambda,t^{\pm 1} v),~(q^{\mp 4}\lambda,t^{\pm 2}v),
{}~\cdots$. Since $q$ is not a root of unity, both the eigenvalues and the
eigenstates are distinct to each other.
This fact would otherwise be contradictory to the finite dimensionality of
$V$ for the finite-dimensional representation $(\pi,V)$, unless
$\lambda=0$. The discussion for $f$ is similar. \\
Thus, {\em if $(\pi,V)$ is a finite-dimensional representation of $SU_q(2)$,
then $e,~f$ act nilpotently on $V$}.
Let $(\pi,V)$ be finite dimension and irreducible, and set $V_{\rm
high}=\{v\in V|ev=0\}$. By the above statement we know that $V_{\rm
high}\not=\{0\}$. Since $tV_{\rm high}=V_{\rm high}$, we can find
such a non-zero vector $v_0\in V_{\rm high}$ that $tv_0=\lambda v_0$
with some nonzero $\lambda$ ($\in {\bf C}^\times$). $f$ also acts
nilpotently, so
that we can find the smallest non-negative integer $l$ satisfying
$$f^jv_0\not=0~~~(0\leq j\leq l),~~~~f^{l+1}v_0=0~.$$
Setting
\begin{equation}
f^{(k)}=\frac{f^k}{[k]!}~,~~~~~~v_k=f^{(k)}v_0~,
\end{equation}
we obtain
\begin{equation}\label{2.chang1}
\begin{array}{rcl}
[e,f^{(k)}]&=&\displaystyle\frac{1}{[k]}[e,f]f^{(k-1)}+
\frac{1}{[k]}f[e,f^{(k-1)}]\\[4mm]
&=&\displaystyle\frac{[k-1]!}{[k]!}[e,f]f^{(k-1)}
+\frac{[k-2]!}{[k]!}f[e,f]f^{(k-2)}\\[4mm]
& & \displaystyle +\frac{[k-2]!}{[k]!}f^2[e,f^{(k-2)}]\\[4mm]
&\vdots&\\[3mm]
&=& \displaystyle\sum_{i=1}^k
\frac{[k-i]!}{[k]!}\left(f\right)^i[e,f]f^{(k-i)}
\\[4mm]
&=&\displaystyle\frac{f^{(k-1)}}{[k](q-q^{-1})}
\sum_{i=1}^k(q^{-2(k-i)}t-q^{2(k-i)}t^{-1})\\[4mm]
&=&f^{(k-1)}\displaystyle\frac{q^{-k+1}t-q^{k-1}t^{-1}}
{q-q^{-1}}~~~~(k\geq 1)~,
\end{array}
\end{equation}
where we have used Eq.(\ref{2.cha1}) repeatedly.\\
Making use of Eq.(\ref{2.chang1}), we have
\begin{equation}
0=[e,f^{(l+1)}]v_0=f^{(l)}\frac{q^{-l}t-q^{l}t^{-1}}{q-q^{-1}}v_0
=\frac{q^{-l}\lambda-q^{l}\lambda^{-1}}{q-q^{-1}}v_l~.
\end{equation}
Since $v_l\not=0$, we get $\lambda=\pm q^l$. Use is made of the
algebra relations, Eq.(\ref{2.cha1}), and the definition for $v_k$, to
get
\begin{equation}
\begin{array}{rcl}
& &tv_k=\pm q^{l-2k}v_k~,\\[2mm]
& &ev_k=[e,f^{(k)}]v_0=\pm[l-k+1]v_{k-1}~,\\[2mm]
& &fv_k=\displaystyle\frac{f^{k+1}}{[k]!}v_0=[k+1]v_{k+1}~,
\end{array}
\end{equation}
where $v_{-1}=0,~v_{l+1}=0$.\\
The above equations can be written in the standard notation as
\begin{equation}\label{2.j1.7}
\begin{array}{rcl}
& &\pi^\sigma_l(e)v_k^l=\sigma[l-k+1]v_{k-1}^l~,\\[2mm]
& &\pi^\sigma_l(f)v_k^l=[k+1]v_{k+1}^l~,\\[2mm]
& &\pi^\sigma_l(t)v_k^l=\sigma q^{l-2k}v_{k}^l~,
\end{array}
\end{equation}
where $l\in Z_{\geq 0}$ and $\sigma=\pm 1$, $v_j^l=0$ if $j>l$ or $j<0$.
Therefore, the representation ($\pi^\sigma_l,V^\sigma(l)$) is a
representation of dimension $(l+1)$.
Obviously $(\pi^+,V^+(l)$) is a $q$-analog to the spin $l/2$ representation of
$su(2)$. However, $(\pi^-,V^-(l)$) has no classical counterpart. This is
attribute to
the presence of an automorphism $\kappa$ of the quantum group $SU_q(2)$ given
by
$$\kappa(e)=-e,~~~ \kappa(f)=f,~~~\kappa(t)=-t~.$$
If $(\pi,V)$ is a representation, one can always ``twist it by sign'' to get
another representation $(\pi\circ\kappa,V)$, e.g.,
$\pi^-_l=\pi^+_l\circ\kappa$. We
shall simply write $(\pi_l^+,V^+(l))$ as $(\pi_l,V(l))$. \\
For the representation $(\pi,V)$, we have
already found a non-zero submodule $V^\sigma(l)=\displaystyle\bigoplus^l_{k=0}
{\bf C}v_k^l\subset V$. Because $V$ is irreducible, we must have
$V^\sigma(l)=V$. \\
Thus {\em an irreducible finite-dimensional representation of
$SU_q(2)$ is equivalent to that of $(\pi_l^\sigma,V^\sigma(l))$.}
If $(\pi,V)$ is a representation, the eigenspace of $t$, $V_\mu=\{v\in
V|tv=q^\mu v\}(\mu\in Z)$ is called as the weight space of weight $\mu$,
and $v\in V_\mu$ is called as a weight vector of weight $\mu$. If $V$ is a
direct sum of weight spaces $\displaystyle\bigoplus_\mu V_\mu$, so is a
submodule $W$: $W=\displaystyle\bigoplus_\mu(W\bigcap
V_\mu)$. In other words, if $w=\displaystyle\sum_\mu
v^{(\mu)},~v^{(\mu)}\in V_\mu$, it is true that $v^{(\mu)}\in W$ for all
$\mu$. In fact, applying $t^k$ we have
\begin{equation}
t^kw=\sum_\mu q^{k\mu}v^{(\mu)}~,~~~~k=0,1,2,\cdots~.
\end{equation}
Treating $v^{(\mu)}$ as unknowns and $t^kw\in W$ as knowns, we can solve
the linear equation set, whose coefficient matrix is a Verdermonde type
$(q^{k\mu})$, and non-singular. This means $v^{(\mu)}\in W$.
Obviously $V=V^\sigma(l)$ is a direct sum of weight spaces
$V_{l-2k}={\bf C}v_k^l$, and $V^\sigma(l)={\bf C}v_0^l\oplus
{\bf C}v_1^l\oplus\cdots\oplus
{\bf C}v_l^l$. Acting $e$ on it, we find that all $v_j^l$ belong to $W$. It
follows that $W=V$. Therefore {\em all representations
$(\pi^\sigma_l,V^\sigma(l))~(l=0,1,2,\cdots,~\sigma=\pm)$ are
irreducible}.
Finally {\em $(\pi_{l}^\pm,V^\pm(l))$ are
inequivalent,} because the sets of eigenvalues of $t$ $\{\pm q^l,\pm
q^{l-2},\cdots,\pm q^{-l}\}$ are distinct to each other.
{}From the above discussion the classification theorem follows.
{\em
\begin{itemize}
\item All representations
$(\pi^\sigma_l,V^\sigma(l))~(l=0,1,2,\cdots,~\sigma=\pm)$ are
irreducible and inequivalent to each other.
\item An irreducible finite-dimensional representation of $SU_q(2)$ is
equivalent to that of $(\pi^\sigma_l,V^\sigma(l))$.
\end{itemize}}
Generally, a representation ($\pi,V)$ with dim$V<\infty$ is reducible,
i.e., there is a submodule $W$ in $V$. Without loss of generality, one may
suppose that the submodule $W$ be
irreducible, and it is the highest weight module with highest weight $l$.
For a reducible $W$, there exists a submodule $W_1\subset W
\subset V$. Repeating this procedure, always one can get an irreducible
submodule $W\subset V$. An irreducible finite-dimensional
representation is characterized by a Casimir operator $C$
\begin{equation}
C=\frac{(qt-1)(1-q^{-1}t^{-1})}{(q-q^{-1})^2}+fe~,
\end{equation}
with the non-zero eigenvalue $\displaystyle\frac{(\sigma
q^{l+1}-1)(1-\sigma^{-1}q^{-(l+1)})}
{(q-q^{-1})^2}$ (here again, $q$ is not a root of unity).
It is easy to verify that the Casimir operator $C$ is an element of the
center of $SU_q(2)$. In other words, $C$ commutes with $e,~f$ and $t^{\pm
1}$. Assuming
\begin{equation}
C'=C-\frac{(\sigma'q-1)(1-(\sigma')^{-1}q^{-1})}{(q-q^{-1})^2}~,
\end{equation}
and acting $C'$ on an irreducible representation, we have
$$\frac{\sigma q^{l+1}+\sigma^{-1}q^{-(l+1)}-\sigma' q-(\sigma')^{-1}q^{-1}}
{(q-q^{-1})^2}~.$$
This eigenvalue is zero, if and only if
$$\frac{q^{l+1}+q^{-(l+1)}}{q+q^{-1}}=\frac{\sigma'}{\sigma}~\in\{1,-1\}~.$$
However, the following equation is not true
$$\frac{q^{l+1}+q^{-(l+1)}}{q+q^{-1}}=1~\longleftrightarrow~
q(q^l-1)=q^{-(l+1)}(q^l-1)~,$$
if the dimension of the representation is greater than $2$
($l\geq 1$ as the dimension of the representation is $(l+1)$).
Similarly, the following equation is not true
$$\frac{q^{l+1}+q^{-(l+1)}}{q+q^{-1}}=-1~\longleftrightarrow~
q(q^l+1)=q^{-(l+1)}(q^l+1)~,$$
if $q$ is not a root of unity. Therefore {\em $C'$ acts in every irreducible
finite-dimensional representation by a scalar, if the dimension of the
representation is greater than $2$}.
If $W$ is codimension $1$ and dim$W\geq 2$, we consider the
representation of $SU_q(2)$ in $V/W$, which is one dimensional: $e,~f$ act
by $0$
and $t$ by a scalar $\sigma$. The acting of $C'$ in $V$ puts $W$ into $W$ by a
nonzero scalar, and in fact, it puts $V$ into $W$ as it acts by $0$ in $V/W$
(by choice of $\sigma'$). Thus there exists such a $1$-dimensional submodule
$W' =$ker$C'$, that $V=W\oplus W'$.
Furthermore, $W'$ is invariant under $SU_q(2)$, because $C'$ belongs to the
center.
If dim$W$=1 and dim$V$=2, the only no-trivial case is for
$\sigma=\sigma'$ with the weight of the representation in $W$ equal to the
weight of the
representation in $W/V$. Thus, there exists a basis $(v_1,v_2)$ in $V$, in
which $t$ has matrix: $\displaystyle\left(\begin{array}{cc}
\sigma&\alpha\\
0 &\sigma\end{array}\right),~~
\alpha\in {\bf C}$,
and
$$t\left(ev_1\right)=\sigma q^2ev_1~,$$
leading to $t(e v_1)=0$;
$$t\left(ev_2\right)= q^2e(\sigma v_2+\alpha v_1)=\sigma q^2
ev_2~,$$
leading to $t(ev_2)=0$ and $e=0$.\\
Similarly, $f=0$. Then the relation $[e,f]=\displaystyle
\frac{t-t^{-1}}{q-q^{-1}}$ implies $t=t^{-1}$, and $\alpha=0$, i.e.,
$t$ is diagonolized. Therefore we obtain $V=W\oplus W'$.
For $W$ with arbitrary codimension, define
\begin{equation}
\begin{array}{rcl}
& &{\cal V}=\{f\in{\cal L}(V,W)\vert f_{|W}~{\rm is~a~scalar~operator}\}~,\\
& &{\cal W}=\{f\in{\cal L}(V,W)\vert f_{|W}=0\}~.
\end{array}
\end{equation}
${\cal W}$ is a submodule of codimension $1$ in ${\cal V}$.
Let $V^*={\rm Hom}(V,C)$ be the dual vector space of $V$. For $v\in V$ and
$v^*\in V^*$, we write $v^*(v)$ as $\langle v^*,v\rangle$. If
$f\in$End$(V)$, $^tf\in$End$(V^*)$ is defined by
$\langle^tf(v^*),v\rangle=\langle v^*,f(v)\rangle$ as usual. Therefore
$^t(f\circ g)=^tg \circ ^tf$. The dual representation
$(\pi^*, V^*)$ is defined by $\pi^*=^t\pi\circ S$:
$$SU_q(2)\stackrel{S}{\to}SU_q(2)\stackrel{^t\pi}{\to}{\rm End}(V^*)~.$$
We make
$SU_q(2)$ act in ${\cal L}(V,W)$ after identifying ${\cal L}(V,W)$
with $W\otimes V^*$ and putting: $\bar{\pi}=(\pi\otimes\pi^*)\circ\Delta$.
For a fixed basis $(y_1,~y_2,~\cdots,~y_p)$ of $W$, we can write
$\phi\in {\cal L}(V,W)$ as $\phi=\sum y_i\otimes x_i^*$ for some
$x_i^*\in V^*$ in a unique manner. ${\cal V}$ and ${\cal W}$ are invariant
under $\bar{\pi}$.
Repeating the arguments for $W$ with codimension $1$, we come to know that
there exists such a submodule ${\cal W'}$ that ${\cal V=W+W'}$. Let
$\phi=\sum y_i\otimes x_i^*$ be a nonzero element in ${\cal W'}$, it acts in
$W$ by a
nonzero scalar and Ker$\phi=\cap_i$Ker$x_i^*$ verifies $V=$Ker$\phi+W$.
Also ,Ker$\phi$ is invariant under $SU_q(2)$ (because ${\cal W'}$ is such
an invariant).\\ The above lead to another important theorem for
the representations of quantum group, the complete reducibility theorem.\\
{\em A finite-dimensional
representation $(\pi,V)$ of $SU_q(2)$ is completely reducible.}\\
This is equivalent to the statement that $V$ is a direct sum of
irreducible
representations. Therefore all finite-dimensional representations of
$SU_q(2)$ are direct sums of some copies of the
$(\pi^\sigma_{l_i},V^\sigma(l_i))$.
A quantum group $A$ is a Hopf algebra. The comultiplication $\Delta$ maps
$A$ into $A\otimes A$ and still satisfies the same algebra relations. As a
result, one may consider tensor products of irreducible finite-dimensional
representations
of $A$. For two representations $(\pi_i,V_i)~(i=1,~2)$, we define the tensor
product representation by the composition of the maps
$$A\stackrel{\Delta}{\to}A\otimes
A\stackrel{\pi_1\otimes\pi_2}{\longrightarrow}
{\rm End}(V_1)\otimes {\rm End}(V_2) \subset {\rm End}(V_1\otimes V_2)~. $$
The tensor product of the two highest weight representations constitutes a new
representation by the action of comultiplication. Generally, the
new representation is reducible. Now, we are discussing the tensor products
for
$SU_q(2)$ in more detail. Provided the two highest weight representations
$$V(m)={\bf C}v_0^m\oplus {\bf C}v_1^m\oplus \cdots\oplus {\bf C}v_m^m~,$$
and
$$V(n)={\bf C}v_0^n\oplus {\bf C}v_1^n\oplus \cdots\oplus {\bf C}v_n^n~,$$
we investigate the irreducible decomposition of $V=V(m)\otimes V(n)$. Let
\begin{equation}
w^l=\sum_{j=0}^s a_jv_j^m\otimes v_{s-j}^n~~~\in V(m)\otimes V(n)~,
\end{equation}
where $s=0,~1,~
\cdots,~{\rm min}(m,n),~l=m+n-2s$. The action of $\Delta(e)$ on $w^l$ yields
\begin{equation}
\Delta(e)(w^l)=\sum_{j=1}^s\left(a_j[m-j+1]+a_{j-1}q^{m-2(j-1)}[n-s+j]\right)
v_{j-1}^m\otimes v_{s-j}^n~,
\end{equation}
and
\begin{equation}
w_0^l=\sum_{j=0}^s a_0(-1)^jq^{j(m+1-j)}\frac{[n-s+j]![m-j]!}{[m]![n-s]!}
v_j^m\otimes v_{s-j}^n
\end{equation}
yields a set of highest weight vectors.\\
The $q$-version of the binomial formula has the form
\begin{equation}\label{zzz}
(A+B)^k=\sum_{i=0}^k q^{i(k-i)}\left[\begin{array}{c}
k\\
i\end{array}\right]A^{k-i}B^i~,
\end{equation}
where $A$ and $B$ are elements in a non-commutative algebra satisfying
$BA=q^2AB$.
Substituting $A=1\otimes f$ and $B=f\otimes t^{-1}$ into Eq.(\ref{zzz}),
yields
\begin{equation}
\begin{array}{rcl}
\Delta\left(f^{(k)}\right)&=&\displaystyle\frac{1}{[k]!}\left(\Delta
(f)\right)^k=\frac{1}{[k]!}\left(
1\otimes f+f\otimes t^{-1}\right)^k\\[3mm]
&=&\displaystyle\sum_{i=0}^k q^{i(k-i)}f^{(i)}\otimes f^{(k-i)}t^{-i}~.
\end{array}
\end{equation}
Acting $\Delta\left(f^{(k)}\right)$ on the set of highest weight vectors
$w_0^l$, we obtain an expression for the general weight vectors $w_k^l$ in
$V(l)\subset V(m)\otimes V(n)$,
\begin{equation}
\begin{array}{rcl}
w_k^l&=&\Delta\left(f^{(k)}\right)w_0^l\\[3mm]
&=&\displaystyle\left(\sum_{i=0}^kq^{i(k-i)}f^{(i)}\otimes
f^{(k-i)}t^{-i}\right)\sum_{j=0}^sa_0
(-1)^jq^{j(m+1-j)}\frac{[n-s+j]![m-j]!}{[m]![n-s]!}
v_j^m\otimes v_{s-j}^n\\[5mm]
&=&\displaystyle\sum_{i=0}^k\sum_{j=0}^sa_0(-1)^jq^{i(k-i)+j(m+1-j)-i(n-2(s-j))}
\frac{[n-s+j]![m-j]!}{[m]![n-s]!}\times\\[4mm]
& &~~~~\times \left[\begin{array}{c}
i+j\\
j\end{array}\right]
\left[\begin{array}{c}
k+s-i-j\\
s-j\end{array}\right]
v_{i+j}^m\otimes v_{s+k-i-j}^n\\[7mm]
&=&\displaystyle\sum_{j=\max(0,k-n+s)}^{\min(m,k+s)}\sum_{\nu=\max(j-k,0)}
^{\min(j,s)}a_0(-1)^\nu q^{\nu(l+1-k)+j(k-j+m-l)}\times\\[4mm]
& &~~~~\times \left[\begin{array}{c}
j\\
\nu\end{array}\right]
\left[\begin{array}{c}
k+s-j\\
s-\nu\end{array}\right]
\left[\begin{array}{c}
n-s+\nu\\
n-s\end{array}\right]
\left[\begin{array}{c}
m\\
\nu\end{array}\right]^{-1} v_j^m\otimes v_{s+k-j}^n~.
\end{array}
\end{equation}
Therefore, we obtain a set of irreducible submodules
$V(m+n),~V(m+n-2),~\cdots,~ V(|m-n|)$ and the Clebsch-Gordan rule
\cite{K88}--\cite{Hou90b}:\\
{\em For any $m,~n\in Z_{\geq 0}$ we have
$V^+(m)\otimes V^+(n)\cong V^+(m+n)\oplus\cdots\oplus V^+(|m-n|)$.}
\subsubsection{Representations of $U_q(g)$, the general case}
It is well-known that the general quantum group $U_q(g)$ satisfies
the following algebra relations
\begin{equation}
\begin{array}{rcl}
& &t_it_i^{-1}=1=t_i^{-1}t_i~,~~~~t_it_j=t_jt_i~,\\
& &t_ie_jt^{-1}_i=q_i^{a_{ij}}e_j~,\\
& &t_if_jt^{-1}_i=q_i^{-a_{ij}}f_j~,\\[2mm]
& &\left[e_i,f_j\right]=\delta_{ij}\displaystyle\frac{t_i-t_i^{-1}}
{q_i-q_i^{-1}}~,\\[3mm]
& &\displaystyle\sum_{\nu=0}^{1-a_{ij}}(-1)^\nu\left[\begin{array}{c}
1-a_{ij}\\
\nu\end{array}\right]
_{q_i}
\left(e_i\right)^{1-a_{ij}-\nu}e_j\left(e_i\right)^\nu=0~,~~~
{\rm for}~i\not= j~,\\[3mm]
& &\displaystyle\sum_{\nu=0}^{1-a_{ij}}(-1)^\nu\left[\begin{array}{c}
1-a_{ij}\\
\nu\end{array}\right]
_{q_i}
\left(f_i\right)^{1-a_{ij}-\nu}f_j\left(f_i\right)^\nu=0~,~~~
{\rm for}~i\not= j~.
\end{array}
\end{equation}
Let $T$ denote the subgroup of invertible elements of $U_q(g)$
generated by $t_i$'s, and $C[T]$ denote its group algebra with basis
$$t_\alpha=t_1^{n_1}t_2^{n_2}\cdots t_l^{n_l},~~~~\alpha=
\sum_{i=1}^ln_i\alpha_i\in Q ~~~(Q=\bigoplus_{i=1}^lZ\alpha_i)~,$$
$U_qn_\pm$ denote the subalgebras generated by $e_i$'s, $f_i$'s with basis
$$\begin{array}{l}
(e)_r=e_1^{m_1} e_2^{m_2} \cdots e_l^{m_l}~,\\
(f)_r=f_1^{m_1} f_2^{m_2} \cdots f_l^{m_l}~,\\
r=\displaystyle\sum_{i=1}^lm_i\alpha_i\in I~~~(I=\bigoplus_{i=1}^lZ_{\geq
0}\alpha_i)~,
\end{array}$$
$U_qb_\pm$ denote the subalgebras generated by $e_i$'s, $t^{\pm 1}_i$'s and
$f_i$'s, $t^{\pm 1}_i$'s with basis
$\left((e)_r\cdot t_\alpha\right)_{r\in I,~\alpha\in Q}$. So,
$U_qb_\pm\simeq U_qn_\pm\otimes C[T]$ as vector spaces. The quantum
group $U_q(g)$ with
basis $\left((e)_r\cdot ( f)_{r'}\cdot t_\alpha\right)
_{r,r'\in I,~\alpha\in Q}$ has a triangular decomposition
$$U_q(g)\simeq U_qn_-\otimes C[T]\otimes U_qn_+$$
as vector spaces and is a free $U_qb_+$-module \cite{Rosso88}.
Let $(\pi,V)$ be representation of $U_q(g)$ in the finite-dimensional
vector space
$V$. Similar to the case of $SU_q(2)$, one may proof that:
{\em
\begin{itemize}
\item The generator $e_i,~f_i$, ($1\leq i\leq l)$ is nilpotent.
\item If the representation is irreducible, the $t_i$'s are diagonalizable
and $V=\bigoplus V_\mu$, where
$$V_\mu=\{v\in V\vert\forall i~t_iv=q_i^{\mu_i}v\}~,~~~~\mu=(\mu_1,~\mu_2,~
\cdots,\mu_l)~,$$
$\mu$'s are the weights of the representation.
\end{itemize}}
Since a 1-dimensional representation is irreducible, we may denote it
by $(\pi_\sigma,{\bf C}_\sigma)$. If $(\pi,V)$ is an irreducible
finite-dimensional representation, $(\pi\otimes\pi_\sigma)\circ
\Delta$ gives an irreducible representation in $V\otimes {\bf C}_\sigma$.
Therefore {\em if $(\pi,V)$ is a irreducible finite-dimensional representation
with highest weight $\mu$, we can associate a
$1$-dimensional representation $(\pi_\sigma,{\bf C}_\sigma)$ and an
irreducible representation with dominant weight
$\tilde{\mu}$ that satisfies $\mu_i\in Z_{\geq 0}$}.\\
A nonzero vector
$v\in V$ is referred as a highest weight vector, if $e_iv=0~~(1\geq i\geq l)$.
For the highest weight vector $v_+$, with weight $\mu=(\mu_1,~\mu_2,~\cdots,~
\mu_l)$, we can construct a cyclic $U_q(g)$-module $V$ spanned by
$$v_0~{\rm and}~f_{i_1}f_{i_2} \cdots f_{i_p} v_0~,~~~~
i_1,~i_2,~\cdots,~i_p\in\{1,~2,~\cdots,~l\}~.$$
$V$ is an indecomposable $U_q(g)$-module, with a unique maximal
proper submodule. Taking the quotient by the maximal proper submodule, we
have an irreducible module with highest weight $\mu:~V_\mu$. For the case that
$\mu$ is of the dominant weight $\tilde{\mu}$, let $w=(f_i)^{\mu_i+1}v_0$.
Using
the algebra relations of $U_q(g)$, we have
$$t_jw=q_j^{-a_{ij}(\mu_i+1)}(f_i)^{\mu_i+1}t_jv_+=q_j^{-a_{ij}(\mu_i+1)}
q_j^{\mu_i}w~.$$
For $w\not= 0$, it is a weight vector with weight ${\mu_i-
(\mu_i+1)a_{ij}}\not=\mu_i$. For $i\not= j,~e_j$ and
$f_i$ commute to each other, and $e_jw=0$. For $i=j$, using the relation
(\ref{2.chang1}) and
$$t_iv_+=q_i^{\mu_i}v_+~,$$
we get $e_iw=0$. As a result, if $w\not= 0$, it would be a highest weight
vector.
Because $V_{\tilde{\mu}}$ is irreducible, such a vector can not exist.
We must have
$$f_i^{\mu_i+1}v_+=0~.$$
For $1\geq i\geq l$, it is easy to verify that the subvectorspace spanned by
$v_0,~f_iv_0,~\cdots,~f_i^{\mu_i}v_0$ is invariant under the action of the
subalgebra $L_i$ generated by $e_i,~f_i$ and $t_i^\pm$. \\
Therefore {\em for each $1\geq i\geq l$, $V_{\tilde{\mu}}$ contains a nonzero
finite-dimensional $L_i$-module.}
For $U_q(g)$, $1-a_{ij}\in\{1,~\cdots,~4\}$, if $1-a_{ij}=1$,
$e_i e_j=e_j e_i$. For $1-a_{ij}\geq 2$, define $e_{i,j}=
e_i e_j-q_i^{a_{ij}}e_j e_i$. If $1-a_{ij}=2$,
one defining relation yields
$$e_i e_{i,j}-q_i^{2+a_{ij}}e_{i,j}e_i=0~.$$
If $1-a_{ij}=3$, put
$$e_{i,i,j}=e_i e_{i,j}-q_i^{2+a_{ij}}e_{i,j} e_i~,$$
and we have
$$e_ie_{i,i,j}=q_i^{4+ a_{ij}}e_{i,i,j}e_i~.$$
For $1-a_{ij}=4$, put
$$e_{i,i,i,j}=e_i e_{i,i,j}-q_i^{4+a_{ij}}e_{i,i,j} e_i~,$$
and then
$$e_ie_{i,i,i,j}=q_i^{6+a_{ij}}e_{i,i,i,j}e_i~.$$
The same relations can be obtained for the $f_i$'s.
Let $V'$ be the sum of the finite-dimensional $L_i$-submodules, obviously
$V'\not=\{0\}$.
If $W$ is an invariant finite-dimensional $L_i$-submodule, the
vector space spanned by $e_j W,~f_j W,~t_jW;~e_{i,j} W,~f_{i,j} W;~
e_{i,i,j}W,~f_{i,i,j}W;$
and $e_{i,i,i,j}W,~f_{i,i,i,j}W,$ (where $j\in\{1,~2,~\cdots,~l\}
\setminus \{i\}$) is finite-dimensional and invariant under $L_i$. So
$U_q(g)(W)\subset V'$. Then we have $V'=V_{\tilde{\mu}}$. \\
We now obtain the basic results about representations of $U_q(g)$ for generic
$q$:
{\em \begin{itemize}
\item If $\mu$ is dominant weight, $V_{\mu}$ is irreducible
and finite-dimensional.
\item An irreducible finite-dimensional representation of $U_q(g)$ is
equivalent to $V_{\tilde{\mu}}$, up to ``twist by sign''.
\end{itemize}}
For the representations of general $U_q(g)$, there is another
important theorem, the complete reducibility theorem.\\
{\em Finite-dimensional representations of $U_q(g)$ are completely
reducible.}
\subsection{Representations for $q$ being a root of unity}
When the quantization parameter $q$ becomes a root of unity ($q^p=\pm 1$),
representation theory
of quantum group changes its phases drastically. This is one of the most
intriguing situations which are absent in the classical case ($q=1$), and
is important also for applications.
We are now discussing the representations for quantum group with the
quantization parameter $q$ is a root of unity. Now the quantum group is not
semi-simple. Let us begin with $SU_q(2)$.
\subsubsection{Representations of $SU_q(2)$}
{\bf The continuity of $V(l)$}
The basic problem for $q^p=\pm 1$ is that
$e^p=0=f^p$, which generates null vectors in some representations.
The representations $(\pi^\sigma_l,V^\sigma(l))$ are well defined also
for $q$ being a root of unity, however they are no longer irreducible in
general. Many representations appearing in the decomposition of tensor
products of irreducible representations will be reducible, but not fully
reducible. It is convenient to construct a different basis for the
representation with highest weight $2j$ \cite{Lusztig90a} as
\begin{equation}
\vert j,m\rangle\equiv v^l_{j-m}~,~~~~(j=\frac{l}{2})~.
\end{equation}
Then, we have
\begin{equation}\label{2.a2}
\begin{array}{l}
e\vert jm\rangle=[j+m+1]\vert j,m+1\rangle~,\\
f\vert jm\rangle=[j-m+1]\vert j,m-1\rangle~,\\[2mm]
\displaystyle\frac{e^a}{[a]!}\vert jm\rangle=\displaystyle
\frac{[j+m+a]!}{[a]![j+m]!}\vert j,m+a\rangle~,\\[2mm]
\displaystyle\frac{f^a}{[a]!}\vert jm\rangle=\displaystyle
\frac{[j-m+a]!}{[a]![j-m]!}\vert j,m-a\rangle~.
\end{array}
\end{equation}
For $q^p=\pm 1$, $[p]=[2p]=\cdots=[kp]=0$. However, the operators
$\displaystyle \frac{e^p}{[p]!}$ and $\displaystyle\frac{f^p}{[p]!}$
are still well defined, as can be
seen by setting $a=p$ in Eq.(\ref{2.a2}) for generic $q$, and then taking the
limit $q^p\to \pm 1$.
It is well-known that the acting of the Casimir operator $C$ on the highest
weight vector $\vert jj\rangle$ yields
\begin{equation}
C\vert jj\rangle=\left[j+\frac{1}{2}\right]^2\vert jj\rangle~.
\end{equation}
For generic $q$ the eigenvalues of $C$ are different for
different values of $j$. For $q^p=\pm 1$, it is easy to see that
the Casimir operator takes identical values for highest weights ${2j}$ and
${2j'}$ related by one of the transformations
\begin{equation}\label{2.tta}
j'=j+np~,~~~~j'=p-1-j+np~.
\end{equation}
The Casimir operator is no longer sufficient to label representations $V(2j)$.
Some $V(2j)$ and $V(2j')$, where $j,~j'$ satisfy Eq.(\ref{2.tta}), can be
mixed up and get connected under the action of $e$.
It is useful to introduce the $q$-dimension
\begin{equation}
D_q(j)=\sum_{{\rm states~ in~}V(2j)}t=[2j+1]~,
\end{equation}
so that
\begin{equation}\label{2.kk1}
D_q(j)=D_q(j-kp)=-D_q(p-1-j+kp)~.
\end{equation}
Both the symmetry properties of the Casimir operator $C$ and the fact that
$e^p=0=f^p$ suggest that if we try to decompose the tensor product
$\left(V(1)\right)^{\otimes n}$ for sufficient high $n$ into irreducible
representations, odd things begin to happen. $\left(V(1)\right)^{\otimes n}$
contains in its decomposition reducible but not fully reducible
representations. For example if $q^3=\pm 1$ we can try to decompose
$V(1)\otimes
V(1)\otimes V(1)$. For generic values of $q$ this tensor product
decomposes into $V(3)\oplus V(1)$. For $q^3=\pm 1$ the weight states
of the $j=\displaystyle\frac{3}{2}$ and one of the
$j=\displaystyle\frac{1}{2}$ representations are mixed up to a reducible,
but not fully reducible representation, because of the presence of null
vectors. It is not difficult to verify that the state
$$
\begin{array}{rcl}
\vert\alpha\rangle&=&\displaystyle f\vert
\frac{1}{2}\frac{1}{2}\rangle\otimes\vert\frac{1}{2}
\frac{1}{2}\rangle\otimes\vert\frac{1}{2}
\frac{1}{2}\rangle \\[2mm]
&=&f\vert v^3_0\rangle
\end{array}$$
is annihilated by
$e$. $\vert \alpha\rangle$ is a highest-weight state of $V(3)$ with
zero norm.
Out of the two other states in $V(1)\otimes V(1)\otimes V(1)$
only one of them is orthogonal to $\vert\alpha\rangle$.
The other state $\vert\beta\rangle$ is not orthogonal to
$\vert\alpha\rangle$, and thus
$$\langle f\alpha\vert\beta\rangle=\langle\alpha\vert e
\vert\beta\rangle\not=0~,$$
i.e., $V(3)$ and $V(1)$ are mixed up (Fig.10). Notice $f^3=0$, implying
that the arrow leaving from $f^2\vert v^3_0\rangle$ to $f^3\vert
v^3_0\rangle$
disappears, while $f^3\vert v^3_0\rangle$ can still be reached from
$\vert v^3_0\rangle$ by applying $\displaystyle\frac{f^3}{[3]!}$. This is
replaced by a new arrow connecting $f\vert\beta\rangle$ and
$\displaystyle\frac{f^3}{[3]!}\vert v^3_0\rangle$. Finally, we obtain
\begin{equation}
ef\vert\beta\rangle=\vert\beta\rangle+f\vert v^3_0\rangle~.
\end{equation}
Thus $(V(1))^3$ decomposes into a big (type-I) representation which is a
mixture of $V(3)$ with $V(1)$, and a small type-II representation $V(1)$.
Because $\displaystyle D_q\left(\frac{3}{2}\right)+D_q\left(\frac{1}{2}\right)
=0$, type-I representation has $q$-dimension zero. The type-I
representation is
indecomposable, but it is not irreducible, because it contains a
sub-representation $V(3)$.
\setlength{\unitlength}{5pt}
\begin{picture}(50,35)(-5,0)
\put(15,3){$f^3\vert v^3_0\rangle$}
\put(20,6){\vector(0,1){5}}
\put(15,12.5){$f^2\vert v^3_0\rangle$}
\put(20,15.5){\vector(0,1){5}}
\put(17,20.5){\vector(0,-1){5}}
\put(10,22){$\vert\alpha\rangle=f\vert v^3_0\rangle$}
\put(17,30){\vector(0,-1){5}}
\put(16,31.5){$\vert v^3_0\rangle$}
\put(32,22){$\vert\beta\rangle$}
\put(35,15.5){\vector(0,1){5}}
\put(32,20.5){\vector(0,-1){5}}
\put(31,12.5){$f\vert \beta\rangle$}
\put(31,25){\vector(-2,1){10}}
\put(31,15.5){\vector(-2,1){10}}
\put(31,11){\vector(-2,-1){10}}
\put(47,22){$\vert\gamma\rangle$}
\put(50,15.5){\vector(0,1){5}}
\put(47,20.5){\vector(0,-1){5}}
\put(46,12.5){$f\vert \gamma\rangle$}
\end{picture}
\hspace{6em}Fig.10. Structure of $(V(1))^{\otimes 3}$ for $q=e^{i\pi/3}$.
The analysis of more general situations can be accomplished in the
same way. In the irrational case, $\left(V(1)\right)^{\otimes n}$ decomposes
into
a sum of representations $V(2j)$. For $q^p=\pm 1$,
we notice that $V(1)$ has positive $q$-dimension. Solving the equation
$D_q(j)=[2j+1]=0$, we find that the $q$-dimension vanishes, whenever
$j=\displaystyle\frac{p-1}{2}+kp$. Making tensor products
$V(1)\otimes V(1)\otimes\cdots\otimes V(1)$ consecutively, we eventually
attain the representation with
$j=\displaystyle\frac{p-1}{2}$ with vanishing $q$-dimension, and others
remained positive $q$-dimension. Further tensoring another copy of $V(1)$
results in pairing of
representations. We can limit the representations of $SU_q(2)$ to those
with the smallest possible $j$ and positive $q$-dimension. This is achieved
by requiring that in the tensor product of the fundamental representations we
only keep
those highest weight vectors that are annihilated by $e$ and not
in the imagine of $e^{p-1}$. This restricts the representations to those
with $j<\displaystyle\frac{p-1}{2}$. In this way we find an alcove in the space
of weights where the $q$-dimension is strictly positive with the lowest
possible
value of $j$. The representations with $j<\displaystyle\frac{p-1}{2}$ are
characterized by the fact that they are highest weight representations and
the highest weight
vector $\vert jj\rangle$ cannot be written as $e^{p-1}\vert{\rm
anything}\rangle$.
For a given values of
the Casimir operator $C$ there are sets of indices like
\begin{equation}
\{j_k>j_{k-1}>\cdots>j_1~,~~0\leq j_1<\frac{1}{2}(p-1)\}~,
\end{equation}
which are related by symmetries (\ref{2.tta}).
If $i$ is odd, $j_i=j_1$ mod $p$, and if $i$ is even, $j_i=p-1-j_1$
mod $p$. If $k=1$, $V(2j_1)$ cannot be mixed up, and it is still an
irreducible highest weight representation (type II). If $k>1$,
mixing of the representations occurs, which is to be analyze.
We construct $V(2j_k)$ by acting $f$ and $\displaystyle\frac{f^p}{[p]!}$
upon the highest weight vector $\vert j_k,j_k\rangle$.
It is easy to see that
\begin{equation}
\begin{array}{rcl}
e\vert j_k,j_{k-1}\rangle&=&[j_k+j_{k-1}+1]\vert j_k,j_{k-1}+1\rangle\\
&=&[np]\vert j_k,j_{k-1}+1\rangle=0~.
\end{array}
\end{equation}
Generally, we have
\begin{equation}
\begin{array}{rcl}
e\vert j_k,j_{k-(2i+1)}\rangle&=&[j_k+j_{k-(2i+1)}+1]\vert j_k,j_{k-(2i+1)}+1
\rangle\\
&=&[n'p]\vert j_k,j_{k-(2i+1)}+1\rangle=0~.
\end{array}
\end{equation}
So that $V(2j_k)$ is not irreducible. It contains states that are annihilated
by $e$ at $j_{k-1},~j_{k-3}=j_{k-1}-p,~j_{k-5}=j_{k-1}-2p,~\cdots.$
It is also easy to see that
\begin{equation}
\begin{array}{rcl}
f\vert j_k,j_{k-2i}+1\rangle&=&[j_k-j_{k-2i}]\vert j_k,j_{k-2i}\rangle\\
&=&[mp]\vert j_k,j_{k-2i}\rangle=0~.
\end{array}
\end{equation}
Then it also contains states that are annihilated by $f$ at
$j_{k-2}+1=j_k-p+1,~j_{k-4}=j_{k}-2p+1,~\cdots.$
The state $\vert j_k,j_{k-1}\rangle $ is annihilated by $e$ and $e^p/[p]!$.
Under the action of $f$ and $f^p/[p]!$ it generates an irreducible
sub-representation, of which all states have zero norm. At spin
$j_{k-1}$, there must exist such a state $\vert \beta\rangle$ that
$e\vert \beta\rangle=\vert j_k,j_{k-1}+1\rangle$. Acting $f$ and
$f^p/[p]!$ upon $\vert\beta\rangle $ generates the states $\vert
j_{k-1},m\rangle$ satisfying,
\begin{equation}\label{2.a16}
f\vert j_{k-1},m\rangle=[j_{k-1}-m+1]\vert j_{k-1},m-1\rangle~.
\end{equation}
{}From the relation
\begin{equation}\label{2.110}
\langle j_k,j_{k-1}+1\vert e\vert j_{k-1},j_{k-1}\rangle=[j_k-j_{k-1}]\langle
j_k,
j_{k-1}\vert j_{k-1},j_{k-1}\rangle~,
\end{equation}
and the normalization
\begin{equation}\label{2.a34}
\langle j_k,j_{k-1}\vert j_{k-1},j_{k-1}\rangle=\frac{1}{[j_{k}-j_{k-1}]}
\left[\begin{array}{c}
2j_{k}\\
j_{k}+j_{k-1}+1
\end{array}\right]~,
\end{equation}
it is follows that
\begin{equation}\label{2.111}
e\vert j_{k-1},j_{k-1}\rangle=\vert j_k,j_{k-1}+1\rangle~.
\end{equation}
In general, the acting of $e$ on $\vert j_{k-1},m\rangle$ has the form
\begin{equation}\label{2.112}
e\vert j_{k-1},m\rangle=[j_{k-1}+m+1]|j_{k-1},m+1>+\displaystyle\left[
\begin{array}{c}
j_k-m-1\\
j_{k-1}-m
\end{array}\right]\vert j_k,m+1\rangle~.
\end{equation}
The above equation can be proved inductively. First, for $m=j_{k-1}$,
Eq.(\ref{2.112}) reduces to Eq.(\ref{2.111}), so that it is satisfied.
Provided that Eq.(\ref{2.112}) is also satisfied for $m=M$, then
\begin{equation}
\begin{array}{rcl}
[2M]\vert j_{k-1},M\rangle
&=&\displaystyle\frac{t-t^{-1}}{q-q^{-1}}\vert j_{k-1},
M\rangle\\[2mm]
&=&\left(ef-fe\right)\vert j_{k-1},M\rangle\\[2mm]
&=&[j_{k-1}-M+1]e\vert j_{k-1},M-1\rangle-[j_{k-1}+M+1][j_{k-1}-M]
\vert j_{k-1},M\rangle\\[4mm]
& &-\left[\begin{array}{c}
j_k-M-1\\
j_{k-1}-M
\end{array}\right][j_k-M]\vert j_k,M\rangle~.
\end{array}
\end{equation}
Thus we obtain
\begin{equation}
e\vert j_{k-1},M-1\rangle=[j_{k-1}+M]\vert j_{k-1},M\rangle +
\left[\begin{array}{c}
j_k-M\\
j_{k-1}-M+1
\end{array}\right]\vert j_k,M\rangle~.
\end{equation}
Therefore Eq.(\ref{2.112}) is also satisfied when $m=M-1$.
Using the normalization (\ref{2.a34}), we have \\ \\
$\langle j_k,-j_{k-1}-1\vert f\vert j_{k-1},-j_{k-1}\rangle$
\begin{equation}
\begin{array}{rcl}
&=&[j_k-j_{k-1}]\langle j_k,-j_{k-1}\vert j_{k-1},-j_{k-1}\rangle\\[2mm]
&=&[j_k-j_{k-1}]\langle j_k,-j_{k-1}\vert \displaystyle
\frac{f^{2j_{k-1}}}{[2j_{k-1}]!}\vert j_{k-1},
j_{k-1}\rangle\\[4mm]
&=&[j_k-j_{k-1}]\left[\begin{array}{c}
j_k+j_{k-1}\\
2j_{k-1}
\end{array}\right]
\langle j_k,j_{k-1}\vert j_{k-1},j_{k-1}\rangle\\[4mm]
&=&\left[\begin{array}{c}
j_k+j_{k-1}\\
2j_{k-1}
\end{array}\right]
\left[\begin{array}{c}
2j_k\\
j_k+j_{k-1}+1
\end{array}\right]~,
\end{array}
\end{equation}
and
\begin{equation}
f\vert j_{k-1},-j_{k-1}\rangle=\left[
\begin{array}{c}
j_k+j_{k-1}-1\\
2j_{k-1}
\end{array}\right]\frac{[j_k+j_{k-1}]}{[j_k-j_{k-1}]}
\vert j_k,-j_{k-1}-1\rangle~.
\end{equation}
{}From the above discussion, we see that $e$ connects $V({2j_k})$ and
$V({2j_{k-1}})$ (Fig.11); $V({2j_{k-1}})$ is the highest weight representation
generated from $\vert \beta\rangle$ if we factorize out $V({2j_{k}})$. From
Eq.(\ref{2.112}), we know that the states in $V(2j_{k-1})$
annihilated by $e$ can be constructed at spin $j_{k-2}=j_k-p$,
$j_{k-4}=j_k-2p$, $\cdots$, by making appropriate linear combinations of
$\vert j_k,m\rangle$ and $\vert j_{k-1},m\rangle $. All states of $(V(2{j_k}),
V(2{j_{k-1}}))$ belonging to ${\rm Ker}~e$ are in ${\rm Im}~e^{p-1}$.
The only states in $(V(2j_k),~V(2j_{k-1}))$ that are annihilated
simultaneously by $e$ and $\displaystyle\frac{e^p}{[p]!}$ are
$\vert \alpha\rangle$ and $\vert \alpha'\rangle=\displaystyle
\frac{f^{j_k-j_{k-1}}}{[j_k-j_{k-1}]!}\vert
\alpha\rangle$, as can be verified by using
formulas (\ref{2.a2}) and (\ref{2.a16}). This fact ensures that there is no
common vector among $(V(2{j_k}),~V(2{j_{k-1}}))$ and its orthogonal.
Therefore, we can repeat the same arguments in the orthogonal of
$(V({2j_k}),V({2j_{k-1}}))$.
\setlength{\unitlength}{4pt}
\begin{picture}(50,120)(-22,0)
\multiput(15,4)(0,1.5){7}{\line(0,1){1}}
\multiput(20,4)(0,1.5){7}{\line(0,1){1}}
\multiput(12.5,14)(0,20){6}{\line(1,0){10}}
\multiput(12.5,19)(0,20){6}{\line(1,0){10}}
\multiput(12.5,24)(0,20){5}{\line(1,0){10}}
\multiput(15,24)(0,1.5){7}{\line(0,1){1}}
\multiput(20,24)(0,1.5){7}{\line(0,1){1}}
\multiput(15,44)(0,1.5){7}{\line(0,1){1}}
\multiput(20,44)(0,1.5){7}{\line(0,1){1}}
\multiput(15,64)(0,1.5){7}{\line(0,1){1}}
\multiput(20,64)(0,1.5){7}{\line(0,1){1}}
\multiput(15,84)(0,1.5){7}{\line(0,1){1}}
\multiput(20,84)(0,1.5){7}{\line(0,1){1}}
\multiput(15,104)(0,1.5){7}{\line(0,1){1}}
\multiput(20,104)(0,1.5){7}{\line(0,1){1}}
\multiput(15,19)(0,20){6}{\vector(0,-1){5}}
\multiput(20,14)(0,20){6}{\vector(0,1){5}}
\multiput(15,24)(0,40){3}{\vector(0,-1){5}}
\multiput(20,39)(0,40){2}{\vector(0,1){5}}
\multiput(45,4)(0,1.5){7}{\line(0,1){1}}
\multiput(50,4)(0,1.5){7}{\line(0,1){1}}
\multiput(42.5,14)(0,20){5}{\line(1,0){10}}
\multiput(42.5,19)(0,20){5}{\line(1,0){10}}
\multiput(42.5,24)(0,20){4}{\line(1,0){10}}
\multiput(45,24)(0,1.5){7}{\line(0,1){1}}
\multiput(50,24)(0,1.5){7}{\line(0,1){1}}
\multiput(45,44)(0,1.5){7}{\line(0,1){1}}
\multiput(50,44)(0,1.5){7}{\line(0,1){1}}
\multiput(45,64)(0,1.5){7}{\line(0,1){1}}
\multiput(50,64)(0,1.5){7}{\line(0,1){1}}
\multiput(45,84)(0,1.5){7}{\line(0,1){1}}
\multiput(50,84)(0,1.5){7}{\line(0,1){1}}
\multiput(45,19)(0,20){5}{\vector(0,-1){5}}
\multiput(50,14)(0,20){5}{\vector(0,1){5}}
\multiput(45,44)(0,40){2}{\vector(0,-1){5}}
\multiput(50,19)(0,40){2}{\vector(0,1){5}}
\multiput(42.5,14)(0,20){5}{\vector(-4,1){20}}
\multiput(42.5,19)(0,20){5}{\vector(-4,1){20}}
\put(-12,18){$j_{k-5}=j_{k-1}-2p$}
\put(-11.5,38){$j_{k-4}=j_{k}-2p$}
\put(-11,58){$j_{k-3}=j_{k-1}-p$}
\put(-10.5,78){$j_{k-2}=j_{k}-p$}
\put(7,118){$j_k$}
\put(6,98){$j_{k-1}$}
\put(16,120){$\vert\alpha\rangle$}
\put(20,100){$\vert\alpha'\rangle$}
\put(46,100){$\vert\beta\rangle$}
\end{picture}
\hspace{7em}Fig.11. Pairing of representations $(V(2j_k),V(2j_{k-1}))$.
In this way we get pairs of
representations mixed with each other in larger structures (type I)
$(V(2{j_k}),V({2j_{k-1}}))$; $(V(2{j_{k-2}}),~V({2j_{k-3}}))$;
$\cdots$. From Eq.(\ref{2.kk1}), it is follows that
\begin{equation}
\begin{array}{ccc}
D_q(j_k)&=&-D_q(j_{k-1})~,\\
D_q(j_{k-2})&=&-D_q(j_{k-3})~,\\
\vdots&\vdots& \vdots~,
\end{array}
\end{equation}
And the type-I representations have $q$-dimension zero.
Depending on the number of $V(1)$'s we end up with a certain
number of $V({2j_1})$ that cannot be mixed up and are still irreducible
highest weight representations (type II).
Type-II representations are described by their highest weight vector
$\vert \alpha_j\rangle$, for which $k\vert \alpha_j\rangle=q^{2j}\vert
\alpha_j \rangle$, $0\leq
j<\displaystyle\frac{1}{2}p-1$. $V(2j)$ is indeed isolated if, moreover
$\vert \alpha_j\rangle$ does not belong to a larger $V(2{j'})$
representation. On basis of
the above analysis, this implies that
$\vert \alpha_j\rangle\in {\rm Im}~e^{p-1}$. Notice that all highest
weights of $V(2{j'})~(j'\geq\displaystyle\frac{1}{2}p-1)$ also belong to
Im $e^{p-1}$. The highest weights of type-II representations are
thus completely characterized by the condition
\begin{equation}\label{2.a19}
\vert \alpha_j\rangle\in\frac{{\rm Ker}~e}{{\rm Im}~e^{p-1}}~.
\end{equation}
Another feature of these states is
\begin{equation}
\vert \alpha_j\rangle\in\frac{{\rm Ker}~f^{2j+1}}{{\rm Im}~f^{p-1-2j}}~.
\end{equation}
For $(V(1))^{\otimes n}$ case, their number reads
\begin{equation}
\Omega_j^{(n)}=\Gamma_j^{(n)}-\Gamma_{p-1-j}^{(n)}+\Gamma_{j+p}^{(n)}
-\Gamma_{p-1-j+p}^{(n)}+\cdots~,
\end{equation}
where
\begin{equation}
\Gamma_j^{(n)}=\left[\begin{array}{c}
n\\
\displaystyle\frac{n}{2}-j
\end{array}\right]-\left[
\begin{array}{c}
n\\
\displaystyle\frac{n}{2}+j+1
\end{array}\right]~.
\end{equation}
A special situation is for
$j_1=\displaystyle\frac{1}{2}(p-1)$, since $p-1-j_1=j_1$, whereupon
the representations $V(2{j_k})$ are still irreducible, and do not pair.
Conventionally, we shall also call them as type I. Then, the representations
space splits into:\\
{\em
\begin{itemize}
\item Type-I representations which have $q$-dimension zero, and are either
mixed up or the kind $V({(np-1)})$.
\item Type-II representations which have a nonzero $q$-dimension, and are
still isomorphic to representations of $U\left(su(2)\right)$.
\end{itemize}}
{\bf Cyclic representations}
It is well-known that the type-II representations $V(2j)~(0\leq j< p/2-1)$
remain irreducible, but have not exhausted the irreducible finite-dimensional
representations, which contain continuous parameters in general. The
continuous parameters arisen from the fact that $SU_q(2)$ has a large
center at roots of unity compared to the generic case.
For definiteness, in dealing with cyclic representations, we
merely consider $q=e^{2\pi i/p}$ with odd $p$
\cite{Concini89}--\cite{Pasquier90}.
Let ${\cal Z}$ denote the center of
$SU_q(2)$
\begin{equation}
{\cal Z}=\{u\in SU_q(2)|au=ua~~~ \forall a\in SU_q(2)\}~.
\end{equation}
Except for the well-known central element, the Casimir operator
\begin{equation} \label{2.mmm}
C=\frac{(qt-1)(1-q^{-1}t^{-1})}{(q-q^{-1})^2}+fe~,
\end{equation}
it is straightforward to show that $e^p,~f^p,~t^p\in{\cal Z}$.\\
Let
\begin{equation}
x=\left((q-q^{-1})e\right)^p~,~~~~ y=\left((q-q^{-1})f\right)^p~,~~~~
z=t^p~,
\end{equation}
and ${\cal Z}_0$ denote the subalgebra of ${\cal Z}$ generated by
$x,~y,~z^\pm$. Then, ${\cal Z}$ {\sl is generated by $C$ and ${\cal Z}_0$.}\\
Rewriting Eq.(\ref{2.mmm}) as
\begin{equation}
fe=C-\frac{(qt-1)(1-q^{-1}t^{-1})}{(q-q^{-1})^2}~,
\end{equation}
we generally have
\begin{equation}
\begin{array}{rcl}
f^pe^p
&=&f^{p-1}\left(fe\right) e^{p-1}\\[3mm]
&=&f^{p-1}e^{p-1}
\left(C-\displaystyle\frac{(q^{2(p-1)+1}t-1)(1-q^{-2(p-1)-1}t^{-1})}{(q-q^{-1})^2}\right)
\\[3mm]
&=&f^{p-2}e^{p-2}
\left(C-\displaystyle\frac{(q^{2(p-2)+1}t-1)(1-q^{-2(p-2)-1}t^{-1})}{(q-q^{-1})^2}\right)\times\\[3mm]
&
&\times\left(C-\displaystyle\frac{(q^{2(p-1)+1}t-1)(1-q^{-2(p-1)-1}t^{-1})}{(q-q^{-1})^2}\right)\\[3mm]
&\vdots&\\[3mm]
&=&\displaystyle\prod_{i=0}^{p-1}\left(C-\frac{(q^{2i+1}t-1)(1-q^{-2i-1}t^{-1})}
{(q-q^{-1})^2}\right)\\[3mm]
&=&\displaystyle\prod_{i=0}^{p-1}\left(C-\frac{(q^it-1)(1-q^{-i}t^{-1})}
{(q-q^{-1})^2}\right)~.
\end{array}
\end{equation}
Introducing the formal operators $u^{\pm 1}$ through the following relation
\begin{equation}
C=\frac{(u-1)(1-u^{-1})}{(q-q^{-1})^2}~,
\end{equation}
we can rewrite $f^p e^p$ as
\begin{equation}
\begin{array}{rcl}
\left((q-q^{-1})f\right)^p\left((q-q^{-1})e\right)^p
&=&\displaystyle\prod_{i=0}^{p-1}\left(u+u^{-1}-q^it-q^{-i}t^{-1}\right)\\
&=&\displaystyle\prod_{i=0}^{p-1}\left(u-q^it\right)\left(1-q^{-i}t^{-1}u^{-1}
\right)~.
\end{array}
\end{equation}
which, upon utilizing the following identities
\begin{equation}
\begin{array}{l}
\displaystyle\prod_{i=0}^{p-1}(A-q^{\pm i}B)=A^p-B^p~~~~(AB=BA)~,\\
\displaystyle\prod_{i=0}^{p-1}(A+q^{\pm i}B)=A^p+B^p~~~~(AB=BA)~,
\end{array}
\end{equation}
reduces to
\begin{equation}
\begin{array}{rcl}
\left((q-q^{-1})f\right)^p\left((q-q^{-1})e\right)^p
&=&\left((\sigma u^p-1)(\sigma -u^{-p})-(\sigma t^p-1)(\sigma -t^{-p})\right)\\
&=&\displaystyle\prod_{i=0}^{p-1}\left(\sigma u-q^i\right)\left(\sigma -
q^{-i}u^{-1}\right)-(\sigma t^p-1)(\sigma -t^{-p})\\
&=&\displaystyle\prod_{i=0}^{p-1}\left(C(q-q^{-1})^2-\sigma (q^i+q^{-i})
+2\right)-(\sigma t^p-1)(\sigma -t^{-p})~.
\end{array}
\end{equation}
This leads to the relation
\begin{equation}\label{2.j4.1}
\psi_p(C)=xy+(\sigma z-1)(\sigma -z^{-1})~,
\end{equation}
where $\psi_p(\mu)=
\displaystyle\prod_{i=0}^{p-1}\left(\mu (q-q^{-1})^2-\sigma (q^i+q^{-i})
+2\right)$. \\
Therefore, {\em the elements $x,~y,~z$ are algebraically independent, while
$C$ is algebraic over ${\cal Z}_0$.}
${\cal Z}_0$ is a Hopf subalgebra, in fact,
\begin{equation}
\Delta(x)=x\otimes 1+z\otimes x~,~~~\Delta(y)=y\otimes z^{-1}+1\otimes y~,~~~
\Delta(z)=z\otimes z~.
\end{equation}
However ${\cal Z}$ is not a Hopf algebra.
Now, let Rep$SU_q(2)$ denote the set of the equivalence classes
of the irreducible finite-dimensional representations
of $SU_q(2)$, Spec${\cal Z}$ denote the hypersurface defined
by Eq.(\ref{2.j4.1}),
$${\rm Spec}{\cal Z}=\{(x,~y,~z,~C\in {\bf C}^4)|z\not=0,~
\psi_p(C)-xy-(\sigma z-1)(\sigma -z^{-1})=0\}~,$$
and further
$${\rm Spec}{\cal Z}_0=\{(x,~y,~z)|z\not=0\}={\bf C}^2\times {\bf
C}^\times~.$$ For an irreducible finite-dimensional representation
$\pi\in{\rm Rep}SU_q(2)$ the central elements $x,~y,~z,~C$
act as scalar in
$\pi$ (Schur's lemma), resulting in the maps
\begin{equation}
{\rm Rep}SU_q(2)\stackrel{X}{\longrightarrow}{\rm Spec}{\cal Z}\stackrel{\tau}
{\longrightarrow}{\rm Spec}{\cal Z}_0~.
\end{equation}
$X$ maps a representation (class) $\pi$ to the values of
$x,~y,~z,~C$ in $\pi$, and the map $\tau$ is a projection, so that
$\tau^{-1}(s)$ consists of $p$ points for general $s\in {\rm Spec}{\cal
Z}_0$. The hypersurface ${\rm Spec}{\cal Z}$ has singularities, given
by the set of $p-1$ points
\begin{equation}
D=\left\{(0,~0,~\sigma,~\frac{\sigma(q^j+q^{-j})-2}{(q-q^{-1})^2}\vert
\sigma=\pm 1,~j=1,~2,~\cdots,~
\frac{p-1}{2}\right\}~.
\end{equation}
De Concini and Kac shown that:
{\em
\begin{itemize}
\item $X$ is surjective.
\item If $\chi \not\in D$, there is only one irreducible finite-dimensional
representation $\pi\in{\rm Rep}
SU_q(2)$ with $X(\pi)=\chi$ and {\rm dim}$\pi=p$.\\
Let $X,~Z$
denote the following $p\times p$ matrices
\begin{equation}
X=\left(\begin{array}{ccccc}
0 & 0 &\cdots& 0&1\\
1 & 0 &\cdots& 0&0\\
0 & 1 &\cdots& 0&0\\
\vdots &\vdots &\ddots&\vdots&\vdots\\
0 & 0 &\cdots&1 &0
\end{array}\right)~,~~~~
Z=\left(\begin{array}{ccccc}
1 & 0 & 0 & \cdots &0\\
0 & q & 0 & \cdots &0\\
0 & 0 & q^2& \cdots &0\\
\vdots&\vdots & \vdots&\ddots &\vdots\\
0 & 0 & 0 &\cdots &q^{p-1}
\end{array}\right)~,
\end{equation}
\begin{equation}
ZX=qXZ~,~~~~Z^p=X^p=1~.
\end{equation}
The $p$-dimensional representation is then given by
\begin{equation}\label{2.jnn}
\begin{array}{rcl}
& &e=\displaystyle\frac{1}{q-q^{-1}}x_1(a_1Z-a_1^{-1}Z^{-1})X~,\\[3mm]
&
&f=\displaystyle\frac{1}{q-q^{-1}}x_1^{-1}(a_2Z^{-1}-a_2^{-1}Z)X^{-1}~,\\[3mm]
& &t=\displaystyle\frac{a_1}{a_2}Z^2~,
\end{array}
\end{equation}
where either $a_1^{2p}\not=1$ or $a_2^{2p}\not=1$, $a_1,~a_2,~x_1\in
{\bf C}^\times$ and
$$x=(a_1^p-a_1^{-p})x_1^p~,~~~y=(a_2^p-a_2^{-p})x_1^{-p}~,~~~
z=\left(\frac{a_1}{a_2}\right)^p~,~~~
C=\frac{\left(qa_1a_2+(qa_1a_2)^{-1}-2\right)}{(q-q^{-1})^2}~,$$
satisfy Eq.(\ref{2.j4.1}).
\item Let $\chi \in D$, then there are exactly two irreducible finite
dimensional representations $\pi^\pm_{j-1},~
\pi_{p-j-1}^\pm\in{\rm Rep}SU_q(2)$ in $X^{-1}(\chi)$, where
$\pi_l^\pm$ signifies the usual highest weight representation
(\ref{2.j1.7}) specialized to $q^p=1$. We have
$${\rm dim}\pi_{j-1}^\pm+{\rm dim}\pi^\pm_{p-j-1}=p~.$$
\end{itemize}}
In conclusion, the independent continuous parameters of the irreducible
finite-dimensional representations of $SU_q(2)$ is three; dim$\pi\leq
p~~(\pi\in {\rm Rep}SU_q(2))$ and the equality holds for general $\pi$.
\subsubsection{Representations of $U_q(g)$, the general case}
{\bf The continuity of regular representations}
For a general quantum group $U_q(g)$, we want to find the
regular representations and the conditions
on the restricted tensor product. Use is made of the Weyl's character
formula to find the regular alcove in
the space of weight. For a representation
with highest weight $\mu$ the expression for the $q$-dimension is
\begin{equation}\label{2.3.39}
D_q(\mu)=\prod_{\alpha>0}\frac{[\langle\mu+\rho,\alpha\rangle]}
{[\langle\rho,\alpha\rangle]}~,
{}~~~~~\rho=\frac{1}{2}\sum_{\alpha>0}\alpha~.
\end{equation}
If $w_i$ is the Weyl reflection with respect
to the simple root $\alpha_i$,
\begin{equation}
\prod_{\alpha>0}\frac{[\langle w_i(\mu)+w_i(\rho), \alpha\rangle]}
{[\langle\rho,\alpha\rangle]}
=-\prod_{\alpha>0}\frac{[\langle\mu+\rho, \alpha \rangle]}{[\langle
\rho,\alpha\rangle]}~,
\end{equation}
and for an element in the Weyl group,
\begin{equation}
\prod_{\alpha>0}\frac{[\langle w(\mu+\rho), \alpha\rangle]}{[\langle
\rho,\alpha\rangle]}
=\epsilon(w)\prod_{\alpha>0}\frac{[\langle\mu+\rho, \alpha\rangle]}
{[\langle\rho,\alpha\rangle]}~,
\end{equation}
where $\epsilon$ is the parity of $w$. Then, we obtain
\begin{equation}
D_q\left(w(\mu+\rho)-\rho+p\sum_{i=1}^ln_i\alpha_i\right)=\epsilon(w)
D_q(\mu)~.
\end{equation}
The $\alpha_i~(i=1,\cdots,l)$ are the simple roots of the classical
algebra $g$. Every positive root can be written as $\alpha=\sum
n_i\alpha_i,~n_i\geq 0$, and $\displaystyle\sum_in_i\equiv{\rm
level}(\alpha)$.
For the highest root $\theta$, the level plus one is
the dual Coxeter number of the algebra
$\hat{g}=(\theta,\theta+2\rho)/\theta^2$. The highest root is normalized
to be length $2$. The largest value of
$(\mu+\rho,\alpha),~\alpha>0$, of a weight $\mu$ is obtained for the highest
root $\theta$, $\langle\rho,\theta\rangle=\hat{g}-1$. The denominator of
Eq.(\ref{2.3.39}) can be written as
\begin{equation}
\prod_{\alpha>0}\left[\langle\rho,\alpha\rangle\right]=
\prod_{l(\alpha)=1}^{\hat{g}-1}
\left[l(\alpha)\right]^{Nl(\alpha)}~,
\end{equation}
where $l(\alpha)$ is the level of $\alpha$ and $Nl(\alpha)$ is the number of
positive roots with the same level. For $q^p=\pm 1,~p>\hat{g}$, the
$q$-dimension
of the generating representations of $U_q(g)$ are positive.
For the representations in increasing values of $\langle\mu,\theta\rangle$
the $q$-dimension remains positive until $\langle\mu,
\theta\rangle=p-\hat{g}$, the $q$-dimension vanishes for
$\langle\mu,\theta\rangle=p-\hat{g}+1$. Beyond this values it can be
positive, negative or zero. And null vectors, reducible but not fully
reducible
representations etc. begin to appear. In analogy with the case of
$SU_q(2)$, let $p=k+\hat{g}$, the regular irreducible representations
acquire the highest weights that are not in
\begin{equation}
{\rm Im}~(e_\theta)^{p-\hat{g}+1}~.
\end{equation}
The first with vanishing $q$-dimension appears as
$\langle\mu,\theta\rangle=k+1$ with $\vert\mu\rangle
=(f_\theta)^{k+1}\vert\alpha\rangle$. Thus, the condition
$\vert\mu\rangle\not\in
{\rm Im}~(e_\theta)^{k+1}$ makes the representations restricted to those
with $\langle\mu,\theta\rangle\leq k$. Because
$\langle\mu,,\theta\rangle\geq\langle\mu,\alpha\rangle,~(\alpha>0)$,
we then obtain $(f_\alpha)^{k+1}\vert\mu\rangle=0,~\forall\alpha.$\\ \\
{\bf Cyclic representations}
Similar to the case of $SU_q(2)$, the new type of the irreducible
finite-dimensional representations appears for general $U_q(g)$, when
$q$ is a root of unity. It is worth finding out the center elements of
$U_q(g)$ first.
To describe the center we need to prepare the braid group actions. For
each $i=1,~\cdots,~l$ the automorphisms $T_i$ of
$U_q(g)$ can be introduced by the following formulas,
\begin{equation}
\begin{array}{rcl}
&&T_i(e_i)=-f_it_i~,\\[3mm]
&&T_i(e_j)=\displaystyle\sum_{m=0}^{-a_{ij}}(-1)^{m-a_{ij}}q_i^{-m}
\frac{e_i^{-a_{ij}-m}}{[-a_{ij}-m]_{q_i}!}e_j
\frac{e_i^{m}}{[m]_{q_i}!}~,~~~{\rm if~}i\not=j~;
\end{array}
\end{equation}
\begin{equation}
\begin{array}{rcl}
&&T_i(f_i)=-t_i^{-1}e_i~,\\[3mm]
&&T_i(f_j)=\displaystyle\sum_{m=0}^{-a_{ij}}(-1)^{m-a_{ij}}q_i^{m}
\frac{f_i^{m}}{[m]_{q_i}!}f_j
\frac{f_i^{-a_{ij}-m}}{[-a_{ij}-m]_{q_i}!}~,~~~{\rm if~}i\not=j~;
\end{array}
\end{equation}
\begin{equation}
T_i(t_j)=t_jt_i^{-a_{ij}}~.~~~
{}~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
\end{equation}
Now, let $w_0$ be the longest element of the Weyl group, and fix a
reduced decomposition
$$w_0=s_{i_1}s_{i_2}\cdots s_{i_\nu}~.$$
It is well-known that the sequence
$$\alpha_{i_1},~s_{i_1}(\alpha_{i_2}),~s_{i_1}s_{i_2}(\alpha_{i_3}),\cdots,
s_{i_1}s_{i_2}\cdots s_{i_\nu}(\alpha_{i_\nu})$$
coincides with the set
of positive roots. For example, if ${ g}= sl(3)$ and
$w_0=s_1s_2s_1$, we have $\alpha_1,~\alpha_1+\alpha_2,~\alpha_2$. This
allows us to define the root vectors $X_{\alpha}$ by
\begin{equation}
\begin{array}{llll}
X_{\alpha_{i_1}}=e_{i_1}~,~~~&X_{\alpha_{i_2}}=T_{i_1}(e_{i_2})~,~~~
&X_{\alpha_{i_3}}=T_{i_1}T_{i_2}(e_{i_3})~,~~~&\cdots~,\\[2mm]
X_{-\alpha_{i_1}}=f_{i_1}~,~~~&X_{-\alpha_{i_2}}=T_{i_1}(f_{i_2})~,~~~
&X_{-\alpha_{i_3}}=T_{i_1}T_{i_2}(f_{i_3})~,~~~&\cdots~.
\end{array}
\end{equation}
As having done for $SU_q(2)$ one can verify
that $e_i^p,~f_i^p,~t_i^p$ belong to the center ${\cal Z}$. Since $T_i$
are automorphisms, $(X_\alpha)^p~(\alpha\in\Delta$, the set of roots) also
belong to ${\cal Z}$. Let Rep$U_q(g)$ denote the equivalence classes of
the irreducible finite-dimensional representations of $U_q(g)$ and
${\cal Z}_0$ denote the subalgebra generated by $(X_\alpha)^p~(\alpha\in
\Delta),~ t_i^{\pm p}~(1\leq i\leq l)$. There exists a natural map
$${\rm Rep}U_q(g)\stackrel{X}{\longrightarrow}{\rm Spec}{\cal Z}~,$$
where Spec${\cal Z}$ means the set of algebra homomorphisms
$\chi:~{\cal Z}\to {\bf C}$.\\
The following theorem can then be proved.
{\em
\begin{itemize}
\item $(X_\alpha)^p~(\alpha\in\Delta),~t_i^p~(1\leq i\leq l)$ are algebraic
independent, and ${\cal Z}$ is algebraic over ${\cal Z}_0$.
\item $X$ is surjective with finite fiber. For general $\chi\in {\rm
Spec}{\cal Z},~X^{-1}(\chi)$ consists of a single representation of
dimension $p^l$.
\end{itemize}}
In particular the theorem indicates that: the number of continuous
parameters
of the irreducible finite-dimensional representations is equal to the
dimension of the classical
algebra $g$. For $\pi\in{\rm Rep}U_q(g)$, the dimension of $\pi$ is less or
equal to $p^l$.
\section{Hamiltonian system}
In this section, through a concrete example -- the symmetric top system,
we discuss quantum symmetry in Hamiltonian systems.
To investigate the motion of rigid body, it is convenient to introduce the
components
of the angular momentum $J_i$ resolved
with respect to the axes in the body frame \cite{Sudarshan,Goldstein}.
The total kinetic energy of the rigid body can be expressed
in terms of $J_i$'s by
\begin{equation}\label{5.1.1}
T=\frac{1}{2}I_{mn}^{-1}J_mJ_n~,
\end{equation}
where $I=||I_{ik}||$ is both symmetric and positive definite, $I_{ik}$ are
components of inertia tensor of the rigid body. In the usual Lagrangian
approach, equations of motion of the rigid body are
\begin{equation}\label{5.1.2}
\dot J_i=\epsilon_{ijm}J_jI_{mn}^{-1}J_n~,
\end{equation}
that possess a simple and elegant form. The interesting point is that the
equations are solely described by $J_i$'s. It is straightforward to
verify that $J_i$'s give the Poisson bracket (PB) realization of the
Lie algebra $su(2)$
\begin{equation}\label{5.1.3}
[J_i,J_j]_{PB}=-\epsilon_{ijk}J_k~.
\end{equation}
If we rewrite the equations of motion in the form
\begin{equation}\label{5.1.4}
\dot J_i=-\epsilon_{mnj}J_j\frac{\partial J_i}{\partial J_m}
\frac{\partial T}{\partial J_n}~,
\end{equation}
they can be put into the standard Hamiltonian formalism
\begin{equation}\label{5.1.5}
\dot J_i=[J_i,T]_{PB}~.
\end{equation}
Because only the variables $J_i$'s appear, the above equations may be directly
computed by just using the Lie algebra $su(2)$ and the derivation property.
Thus, as far as the equations of motion are concerned, all that we need are
the expression of the Hamiltonian as a function of the $J_i$'s, and the
Poisson brackets among the $J_i$'s. $J_i$'s do not form a complete set of
dynamical variables of rigid body rotation. However, there is a family of
classical systems, each of which is completely described by the variables
$J_i$'s, with proper physical interpretation. The
equations of motion can thus be written in the Hamiltonian form but using a
generalized Poisson bracket (GPB). The basic GPB's among $J_i$'s are
postulated to have the form
\begin{equation}\label{5.1.6}
[J_i,J_j]_{GPB}=-\epsilon_{ijk}J_k~,
\end{equation}
and the GPB of any two functions $f({\bf J})$ and $g({\bf J})$ is computed
by the derivation property
\begin{equation}\label{5.1.7}
[f({\bf J}),g({\bf J})]_{GPB}=-\epsilon_{mni}J_i
\frac{\partial f}{\partial J_m}
\frac{\partial g}{\partial J_n}~.
\end{equation}
Such classical dynamical systems are called as the classical pure-spin systems,
which differ from the systems of rotating rigid body:
\begin{itemize}
\item for the former, all dynamical variables are defined as suitable functions
of the $J_i$'s;
\item the three-dimensional space with the $J_i$'s as coordinates, each with
a specified range, forms the generalized phase space of the systems;
\item the solution of the equations of motion that equate each
time
derivative $\dot J_i$ with the GPB of $J_i$'s with a Hamiltonian
$H({\bf J})$, amounts to a complete solution of the motion and suffices
to determine the ''phase'' of the system at
any time in terms of its ''phase'' at an earlier time.
\end{itemize}
For a pure-spin system, because the $J_i$'s form a complete set of
dynamical variables, there are no other invariants than the Casimir's,
and there is only one of them, to wit, $J^2=\displaystyle\sum_{i=1}^3
J_iJ_i$. This Casimir is invariant
under all generalized canonical transformations. It thus follows that each
transformation of such kind
acts as a canonical mapping of surface of sphere onto themselves
in the three-dimensional space of $J_i$'s, which preserving the radii and the
basic
GPB's. Therefore, the phase space for a pure-spin system is composed of some
set of
surfaces of spheres centered on the origin in the three-dimensional space.
For example, it could be the entire three-dimensional space, with each $J_i$'s
varying independently from $-\infty$ to $+\infty$, or it could be only the
surface of some sphere, with $J^2$ having some given numerical value
characteristic of the system and with only two independent $J_i$.
Various intermediate possibilities are easily
conceivable. To preserve the symmetry of the formulae, even if $J^2$ is
constrained to have some given value, we treat all three $J_i$ as independent
variables for partial differentiation. As long as the
partial differentiation are merely associated with the computations of
GPB's, the value of $J^2$ follows at the end of all
calculations.
\subsection{Classical symmetric top system}
The standard Hamiltonian of a symmetric top is
\begin{equation}\label{5.e233}
H=\frac{J_1^2+J_2^2}{2I}+\frac{J_3^2}{2I_3}~.
\end{equation}
By the above discussion, the phase space of the symmetric top
system can be of the form
\begin{equation}\label{5.e197}
M_0:~~~~~~J_{1}^{2}+J_{2}^{2}+J_{3}^{2}=J_{0}^{2}~.
\end{equation}
In the phase space $M_0$, we define a symplectic form by
\begin{equation}\label{5.e235}
\Omega_{0}=\displaystyle\frac{1}{2J_0^{2}}\sum_{i,j,k=1}^3 \epsilon_{ijk}
J_{i} dJ_{j}\wedge dJ_{k}~.
\end{equation}
For the one-parameter group of the rotations around the $i$-th axis there
exist Hamiltonian vector fields $X_{J_i}$ with respect to $\Omega_0$
\begin{equation}\label{5.1.11}
X_{J_i}=\sum_{j,k=1}^3\epsilon_{ijk}J_j\frac{\partial}{\partial J_k}~.
\end{equation}
By using the relation (\ref{5.e197}), it is not difficult to show that the
following relations are satisfied
\begin{equation}\label{5.1.12}
X_{J_i}\rfloor\Omega_0=-dJ_i,~~~~~~~\Omega_0(X_{J_i},X_{J_j})
=[J_i,J_j]_{GPB}~.
\end{equation}
The generalized Poisson brackets among the variables $J_{i}$ read
\begin{equation}\label{5.2.8}
[J_i,J_j]_{GPB}=-X_{J_i} J_{j}=-\epsilon_{ijk} J_{k}~,
\end{equation}
or
\begin{equation}\label{5.e202}
[J_{3},J_{\pm}]_{GPB}=\pm i J_{\pm}~,~~~
[J_{+},J_{-}]_{GPB}=i2J_{3}~,
\end{equation}
where $J_\pm=J_1\pm iJ_2$.
{}From the above Lie algebra $su(2)$, it is not difficult to show that
\begin{equation}\label{5.2.10}
\begin{array}{l}
\dot J_1=[J_1,H]_{GPB}=\displaystyle\frac{I-I_3}{II_3}J_2J_3~,\\[3mm]
\dot J_2=[J_2,H]_{GPB}=\displaystyle\frac{I_3-I}{II_3}J_1J_3~,\\[3mm]
\dot J_3=[J_3,H]_{GPB}=0~.
\end{array}
\end{equation}
Thus, we obtained the equations of motion for the symmetric top in terms of
the symplectic geometry.
Following the above method, we are going to discuss the symmetric
top system with deformed Hamiltonian and symplectic structure
\cite{Chang92b}--\cite{Chang92a}. To begin with, we write the deformed
Hamiltonian of the
symmetric top as,
\begin{equation}\label{5.2.31}
H_q=\frac{I-I_3}{2II_3}J_3^{\prime 2}+\frac{1}{2I}\left(J_1^{\prime 2}+
J_2^{\prime 2}+
\frac{(\sinh\gamma J'_3)^{2}}{\gamma \sinh\gamma}\right)~.
\end{equation}
The observables $J_i^\prime$'s are related with $J_i$'s by
\begin{equation}\label{5.F1}
\begin{array}{l}
J_1^\prime=\displaystyle\frac{1}{\sqrt{\gamma\sinh\gamma}}
\sqrt{\displaystyle\frac{\sinh\gamma(J_0+J_3)\sinh\gamma(J_0-J_3)}
{(J_0+J_3)(J_0-J_3)}}J_1~,\\[4mm]
J_2^\prime=\displaystyle\frac{1}{\sqrt{\gamma\sinh\gamma}}
\sqrt{\displaystyle\frac{\sinh\gamma(J_0+J_3)\sinh\gamma(J_0-J_3)}
{(J_0+J_3)(J_0-J_3)}}J_2~,\\[4mm]
J_3^\prime=J_3~.
\end{array}
\end{equation}
In terms of the $J_i^\prime$'s, Eq.(\ref{5.e197}) is
\begin{equation}\label{5.e205}
M_0^q:~~~~~J_{1}^{\prime 2}+J_{2}^{\prime 2}+\displaystyle
\frac{(\sinh\gamma J_{3}^\prime)^{2}}
{\gamma \sinh\gamma}=J_q^{2}~,
\end{equation}
where $J_q=\displaystyle\frac{\sinh\gamma J_{0}}{\sqrt{\gamma\sinh\gamma}}$.
On the deformed sphere $M_0^q$, the symplectic form \cite{Fei91,Fei92} is
defined by
\begin{equation}\label{5.e206}
\Omega_q=\displaystyle\frac{1}{J^{2}_q}
(J_1^\prime dJ_2^\prime \wedge dJ_3^\prime +J_2^\prime dJ_3^\prime
\wedge dJ_1^\prime
+\displaystyle\frac{\tanh\gamma J_3^\prime}{\gamma} dJ_1^\prime
\wedge dJ_2^\prime)~.
\end{equation}
The Hamiltonian vector fields $X_{J_{i}^\prime}$ on $M_0^q$
now possess the form
\begin{equation}
\begin{array}{l}
X_{J_1^\prime}=-J_2^\prime \displaystyle\frac{\partial}{\partial J_3^\prime}+
\displaystyle\displaystyle\frac{\sinh 2\gamma J_3^\prime}
{2\sinh\gamma}\displaystyle\frac{\partial}{\partial J_2^\prime}~,\\[6mm]
X_{J_2^\prime}=-\displaystyle\frac{\sinh 2 \gamma J_3^\prime}
{2 \sinh \gamma}\displaystyle\frac{\partial}{\partial
J_1^\prime}+J_1^\prime \displaystyle\frac{\partial}{\partial
J_3^\prime}~,\\[6mm]
X_{J_3^\prime}=-J_1^\prime\displaystyle\frac{\partial}{\partial J_2^\prime}
+J_2^\prime\displaystyle\frac{\partial}{\partial J_1^\prime}~.
\end{array}
\end{equation}
It is not difficult to verify that the Hamiltonian vector fields satisfy the
relations
\begin{equation}
\begin{array}{l}
X_{J_i^\prime}\rfloor \Omega_q=-dJ_i^\prime~,\\[4mm]
[X_{J_i^\prime},X_{J_j^\prime}]=-X_{[J_i^\prime,J_j^\prime]_{GPB}}~,\\[4mm]
\Omega_q(X_{J_i^\prime},X_{J_j^\prime})=[J_i^\prime,J_j^\prime]_{GPB}~.
\end{array}
\end{equation}
Then, we get the basic generalized Poisson brackets,
\begin{equation}\label{5.e209}
\begin{array}{l}
[J_1^\prime,J_2^\prime]_{GPB}=-\displaystyle\displaystyle\frac
{\sinh 2 \gamma J_3^\prime}{2\sinh\gamma}~,\\[4mm]
[J_2^\prime,J_3^\prime]_{GPB}=-J_1^\prime~,
{}~~~~~~[J_3^\prime,J_1^\prime]_{GPB}=-J_2^\prime~;
\end{array}
\end{equation}
or
\begin{equation}
[J_+^\prime,J_-^\prime]_{GPB}=i \displaystyle
\frac{\sinh 2\gamma J_3^\prime}{\sinh\gamma}~,~~~
[J_3^\prime,J_{\pm}^\prime]_{GPB}=\pm i J_{\pm}^\prime~.
\end{equation}
This algebra is the quantum algebra $SU_{q,\hbar\rightarrow 0}(2)$
\cite{Chang90a}--\cite{Flato}.
We now introduce two open sets $U_\pm$ on the phase space $M_0^q$,
\begin{equation}
U_{\pm}=\left\{ x \in M_0^q \mid J_q \pm
\displaystyle\frac{\sinh \gamma J_3^\prime}
{\sqrt{\gamma \sinh \gamma}}\not= 0\right\}~,
\end{equation}
and two complex functions $ z_+$ and $ z_-$ on $ U_+ $ and $ U_-$,
respectively,
\begin{equation}\label{5.e212}
z_{\pm}=(J_1^\prime\mp i J_2^\prime)\left(J_q\pm\displaystyle
\frac{\sinh\gamma J_3^\prime}
{\sqrt{\gamma \sinh\gamma}}\right)^{-1}~.
\end{equation}
In $U_+\bigcap U_-$ we have
\begin{equation}
z_{+} z_{-} =1~.
\end{equation}
In order to construct a Hopf algebra structure for
the quantum algebra $SU_{q,\hbar\rightarrow 0}(2)$, we first search for a set
of classical
operators $\tilde{J}_{i}^\prime$ which give the Lie bracket realization of the
quantum algebra by using the pre-quantization method \cite{Sniatychi}.
{}From the definition of complex coordinates $z_{+}$ and
$z_{-}$ introduced on $M_0^q$ of Eq.(\ref{5.e212}), we get the expressions for
$J_{i}^\prime$'s, in terms of $z_{+}$ and $z_{-}$,
\begin{equation}\label{5.e214}
\begin{array}{l}
J_1^\prime=-J_q\displaystyle\displaystyle\frac{z_{\pm}+\overline{z}_{\pm}}
{1+z_{\pm}\overline{z}_{\pm}}~,~~~
J_2^\prime=\mp i J_q\displaystyle\displaystyle\frac{z_{\pm}-\overline{z}_{\pm}}
{1+z_{\pm}\overline{z}_{\pm}}~,\\[5mm]
\displaystyle\frac{\sinh\gamma J_3^\prime}{\sqrt{\gamma \sinh\gamma}}
=\mp J_q\displaystyle\frac{1-
z_{\pm}\overline{z}_{\pm}}{1+z_{\pm}\overline{z}_{\pm}}~.
\end{array}
\end{equation}
Then, we can write the q-deformed symplectic
form (\ref{5.e206}) as
\begin{equation}
\begin{array}{rcl}
\Omega_q\vert_{ U_{\pm}}&=&2 i J_q\left(J^{2}_q\gamma^{2}
\displaystyle\frac{(1-
z_{\pm}\overline{z}_{\pm})^2}{(1+z_{\pm}\overline{z}_{\pm})^{2}}+
\frac{\gamma}{\sinh\gamma}\right)^{-\frac{1}{2}}
\displaystyle\frac{d\overline{z}_{\pm}\wedge dz_{\pm}}
{(1+z_{\pm}\overline{z}_{\pm})^2}\\[6mm]
&=&-i Q_{\pm}d\overline{z}_{\pm}\wedge dz_{\pm}~,
\end{array}
\end{equation}
where
\begin{equation}
Q_{\pm}=-2 J_q
\left(J^{2}_q\gamma^{2}\displaystyle\frac{(1-z_{\pm}\overline{z}_{\pm})^2}
{(1+z_{\pm}\overline{z}_{\pm})^{2}}+
\frac{\gamma}{\sinh\gamma}\right)^{-\frac{1}{2}}(1+z_{\pm}\overline{z}_{\pm})^{-2}~.
\end{equation}
Since $\Omega_q$ is closed, it should be locally exact on the open set
$U_+$ and $U_-$, i.e.,
\begin{equation}\label{5.2.45}
\Omega_q\vert_{U_{\pm}}=d \theta_{\pm}~.
\end{equation}
Here the symplectic one forms $ \theta_{\pm}$ read
\begin{equation}\label{5.2.46}
\begin{array}{rcl}
\theta_{\pm}&=&-\displaystyle\frac{i}{\gamma z_{\pm}}\left(\sinh^{-1} \left(J_q
\sqrt{\gamma \sinh\gamma}\frac{1-z_{\pm}\overline{z}_{\pm}}
{1+z_{\pm}\overline{z}_{\pm}}\right)-\sinh^{-1}\left(J_q
\sqrt{\gamma \sinh\gamma}\right)\right)dz_{\pm}\\[4mm]
&=&-i p_{\pm}dz_{\pm}~,
\end{array}
\end{equation}
where
\begin{equation}\label{5.2.47}
p_{\pm}=\displaystyle\frac{1}{\gamma z_{\pm}}\left(\sinh^{-1}\left(J_q
\sqrt{\gamma \sinh\gamma}\displaystyle\frac{1-z_{\pm}\overline{z}_{\pm}}
{1+z_{\pm}\overline{z}_{\pm}}\right)-\sinh^{-1}\left(J_q
\sqrt{\gamma \sinh\gamma}\right)\right)~.
\end{equation}
The Hamiltonian vector fields of $J_{i}^\prime$'s now possess the form
\begin{equation}\label{5.e220}
\begin{array}{l}
X_{J_1^\prime}=i \sqrt{\displaystyle\displaystyle\frac{\gamma}{\sinh\gamma}}
\displaystyle\displaystyle\frac{\cosh\gamma J_3^\prime}{2}
\left((z_{\pm}^{2}-1)\displaystyle\displaystyle\frac{\partial}
{\partial z_{\pm}}+(1-\overline{z}_{\pm}^{2})
\displaystyle\displaystyle\frac{\partial}{\partial \overline{z}_{\pm}}\right)~,
\\[6mm]
X_{J_2^\prime}=\mp\sqrt{\displaystyle\displaystyle\frac{\gamma}{\sinh\gamma}}
\displaystyle\displaystyle\frac{\cosh\gamma J_3^\prime}{2}
\left((z_{\pm}^{2}+1)\displaystyle\displaystyle\frac{\partial}{\partial
z_{\pm}}+
(1+\overline{z}_{\pm}^{2})
\displaystyle\displaystyle\frac{\partial}{\partial \overline{z}_{\pm}}\right)~,
\\[6mm]
X_{J_3^\prime}=\mp i
\left(\overline{z}_{\pm}\displaystyle\displaystyle\frac{\partial}
{\partial \overline{z}_{\pm}}-z_{\pm}\displaystyle\displaystyle\frac{\partial}
{\partial \overline{z}_{\pm}}\right)~.
\end{array}
\end{equation}
Let us rewrite expressions (\ref{5.e214}) in terms of the variables $z_{\pm}$
and $p_{\pm}$,
\begin{equation}\label{5.e221}
\begin{array}{rcl}
J_1^\prime&=&-\displaystyle\frac{1}{\sqrt{\gamma\sinh\gamma}}
\left(\cosh\left(\frac{\gamma}{2}z_{\pm}p_{\pm}\right)
z_{\pm}\sinh\left(\frac{\gamma}{2}\left(z_{\pm}p_{\pm}+2b\right)
\right)\right.\\[4mm]
&&~~~~~~\displaystyle\left.-\cosh\left(\frac{\gamma}{2}\left(z_{\pm}p_{\pm}+2b\right)
\right)\frac{1}{z_{\pm}}\sinh\left(\frac{\gamma}{2}z_{\pm}p_{\pm}\right)
\right)~,\\[4mm]
J_2^\prime&=&\displaystyle\frac{\mp i}{\sqrt{\gamma\sinh\gamma}}
\left(\cosh\left(\frac{\gamma}{2}z_{\pm}p_{\pm}\right)
z_{\pm}\sinh\left(\frac{\gamma}{2}\left(z_{\pm}p_{\pm}+2b\right)
\right)\right.\\[4mm]
&&~~~~~~\displaystyle+\left.\cosh\left(\frac{\gamma}{2}\left(z_{\pm}p_{\pm}+2b\right)
\right)\frac{1}{z_{\pm}}\sinh\left(\frac{\gamma}{2}z_{\pm}p_{\pm}
\right)\right)~,\\[4mm]
J_3^\prime&=&\mp\left(z_{\pm}p_{\pm}+b\right)~.
\end{array}
\end{equation}
Here, for convenience, we have used the relation
\begin{equation}\label{5.e267}
\sinh\gamma b=J_q\sqrt{\gamma\sinh\gamma}~.
\end{equation}
The Hamiltonian vector fields of $z$ and $p$ are
\begin{equation}\label{5.e268}
X_{z_{\pm}}=-i Q_{\pm}^{-1}\displaystyle\displaystyle\frac{\partial}
{\partial \overline{z}_{\pm}}~,~~~
X_{p_{\pm}}=i\displaystyle \displaystyle\frac{\partial}{\partial z_{\pm}}-
i Q_{\pm}^{-1}\displaystyle\frac{\partial p_{\pm}}{\partial z_{\pm}}
\displaystyle\displaystyle\frac{\partial}{\partial \overline{z}_{\pm}}~,
\end{equation}
where the relation $\displaystyle\frac{\partial p_{\pm}}{\partial\overline{z}_
{\pm}}=Q_{\pm}$ has been used.\\
We get their pre-quantization operator representations as,
\begin{equation}\label{5.e224}
\tilde{z}_{\pm}=-Q_{\pm}^{-1}\displaystyle\displaystyle\frac{\partial}
{\partial \overline{z}_{\pm}}-z_{\pm}~,~~~
\tilde{p}_{\pm}=\displaystyle\displaystyle\frac{\partial}{\partial z_{\pm}}-
Q_{\pm}^{-1}\displaystyle\frac{\partial p_{\pm}}{\partial z_{\pm}}
\displaystyle\displaystyle\frac{\partial}{\partial \overline{z}_{\pm}}~.
\end{equation}
It is not difficult to verify that the commutator of $ \tilde{z}$ and $
\tilde{p}$ is
\begin{equation}\label{5.e225}
[\tilde{z}_{\pm},\tilde{p}_{\pm}]=1~.
\end{equation}
By means of the formulas of Eq.(\ref{5.e224}), the pre-quantization operators
with
respect to Eq.(\ref{5.e221}) can be expressed as
\begin{equation}\label{5.e226}
\begin{array}{rcl}
\tilde{J}_{1}^\prime&=&-\displaystyle\frac{1}{\sqrt{\gamma\sinh\gamma}}
\left(\cosh\left(\frac{\gamma}{2}\tilde{z}_{\pm}\tilde{p}_{\pm}\right)
\tilde{z}_{\pm}\sinh\left(\frac{\gamma}{2}\left(\tilde{z}_{\pm}
\tilde{p}_{\pm}+2b\right)\right)\right.\\[4mm]
&&~~~~~~\left.\displaystyle-\cosh\left(\frac{\gamma}{2}\left(\tilde{z}_{\pm}
\tilde{p}_{\pm}+2b\right)\right)\frac{\gamma\tilde{p}_{\pm}}{2}
\sum_{n=0}^{\infty}\left(\left(2n+1\right)!\right)^{-1}\left(
\frac{\gamma}{2}\tilde{z}_{\pm}\tilde{p}_{\pm}\right)^{2n}\right)~,\\[4mm]
\tilde{J}_{2}^\prime&=&\displaystyle\frac{\mp i}{\sqrt{\gamma\sinh\gamma}}
\left(\cosh\left(\frac{\gamma}{2}\tilde{z}_{\pm}\tilde{p}_{\pm}\right)
\tilde{z}_{\pm}\sinh\left(\frac{\gamma}{2}\left(\tilde{z}_{\pm}
\tilde{p}_{\pm}+2b\right)\right)\right.\\[4mm]
&&~~~~~~\displaystyle\left.+\cosh\left(\frac{\gamma}{2}\left(\tilde{z}_{\pm}
\tilde{p}_{\pm}+2b\right)\right)\frac{\gamma\tilde{p}_{\pm}}{2}
\sum_{n=0}^{\infty}\left(\left(2n+1\right)!\right)^{-1}\left(
\frac{\gamma}{2}\tilde{z}_{\pm}\tilde{p}_{\pm}\right)^{2n}\right)~,\\[4mm]
\tilde{J}_{3}^\prime&=&\mp\left(\tilde{z}_{\pm}\tilde{p}_{\pm}+b\right)~.
\end{array}
\end{equation}
It is straightforward to verify that they yield the Lie bracket realization of
the quantum algebra $SU_{q,\hbar\rightarrow 0}(2)$ by owing to
Eq.(\ref{5.e225}),
\begin{equation}\label{5.e227}
[\tilde{J}_{1}^\prime,\tilde{J}_{2}^\prime]= -i\displaystyle
\displaystyle\frac{\sinh 2 \gamma\tilde{ J}_{3}^\prime}
{2 \gamma}~,~~~~
[\tilde{J}_{2}^\prime,\tilde{J}_{3}^\prime ]= -i \tilde{J}_{1}^\prime~,~~~~
[\tilde{J}_{3}^\prime,\tilde{J}_{1}^\prime ]= -i \tilde{J}_{2}^\prime~,
\end{equation}
or
\begin{equation}\label{5.e228}
[\tilde{J}_{+}^\prime,\tilde{J}_{-}^\prime]=-\displaystyle
\frac{\sinh 2 \gamma \tilde{J}_{3}^\prime}
{\gamma}~,~~~
[\tilde{J}_{3}^\prime,\tilde{J}_{\pm}^\prime]=\mp \tilde{J}_{\pm}^\prime~.
\end{equation}
Keeping the operators of Eq.(\ref{5.e226}) and commutators of Eq.(\ref{5.e227})
in mind,
we can define the Hopf algebra structure of the classically
realized quantum algebra $SU_{q,\hbar\rightarrow 0}(2)$ as follows
\begin{equation}
\begin{array}{l}
\Delta(\tilde{J}_3')=\tilde{J}_3'\otimes 1+1\otimes \tilde{J}_3' ~,\\
\Delta(\tilde{J}_\pm')=\tilde{J}_\pm'\otimes e^{-\gamma\tilde{J}_3'}
+e^{\gamma\tilde{J}_3'}
\otimes \tilde{J}_3' ~,\\
\epsilon(\tilde{J}_\pm') =\epsilon(\tilde{J}_3') =0~,\\
S(\tilde{J}_\pm')=-q^{\pm 1}\tilde{J}_\pm'~,~~~~S(\tilde{J}_3')=-
\tilde{J}_3' ~.
\end{array}
\end{equation}
It is straightforward to calculate the equations of motion for the classical
symmetric top system with deformed Hamiltonian and deformed symplectic
structure
by using the quantum group $SU_{q,\hbar\to 0}(2)$
\begin{equation}\label{5.2.60}
\begin{array}{rcl}
\dot J_1^\prime&=&[J_1^\prime,H_q]_{GPB}=\displaystyle\frac{I-I_3}{II_3}
J^\prime_2J^\prime_3~,\\[3mm]
\dot J_2^\prime&=&[J^\prime_2,H_q]_{GPB}=\displaystyle\frac{I_3-I}{II_3}
J^\prime_1J^\prime_3~,\\[3mm]
\dot J^\prime_3&=&[J^\prime_3,H_q]_{GPB}=0~.
\end{array}
\end{equation}
Because the deformed Hamiltonian system $(M^q_0,\Omega_q,H_q)$
obeys the same equations of motion with the standard Hamiltonian system
$(M_0,\Omega_0,H)$, the Hamiltonian systems $(M_0,\Omega_0,H)$ and
$(M_0^q,\Omega_q,H_q)$
describe the same physical motions.
It is worth noting that the symmetry of the standard Hamiltonian system
$(M_0,\Omega_0,H)$ is the Lie group $SU(2)$, however,
the symmetry possessed by the deformed Hamiltonian system
$(M_0^q,\Omega_q,H_q)$ is the quantum group $SU_{q,\hbar\to 0}(2)$.
It is well-known that the equations of motion for the standard symmetric top
can be
solved exactly. Therefore the deformed Hamiltonian system
$(M_0^q,\Omega_q,H_q)$
is the model that can be solved exactly.
\subsection{Quantum symmetric top system}
By means of geometric quantization, we now discuss quantum symmetry in quantum
symmetric top system. Let us begin with writing down the deformed
Hamiltonian of the symmetric top \cite{Chang92c,Chang92a}
\begin{equation}\label{5.ssss}
\hat{H_q}=\frac{\hbar^2(I-I_3)}{2II_3}J_3^{\prime 2}+\frac{\hbar^2}{2I}
\left(\frac{\gamma}{\sinh\gamma}\hat{J}_{+}^\prime \hat{J}_{-}^\prime
+[\hat{J}_{3}^\prime ][\hat{J}_{3}^\prime +1]\right)~.
\end{equation}
The geometric quantization
of the deformed Hamiltonian system $(M_0^q,\Omega_q,H_q)$ is described by the
pre-quantization line bundle $L_q$ and the polarization $F$ \cite{Sniatychi}.
For the case under consideration such a
quantum line bundle $L_{q}$ exists if and only if $(2\pi)^{-1}\Omega_q$
defines an integral de Rham cohomology class, i.e., the de Rham cohomology
class $\{-(2\pi)^{-1}\Omega_q\}$ of $-(2\pi)^{-1}\Omega_q$ should be
integrable. Integrating the right hand side of Eq.(\ref{5.e206}) over the
symplectic manifold $M_0^q$, we have
\begin{equation}
\int_{M_0^q}\Omega_q=\displaystyle\frac{1}{J_q^{2}}\left.\left(2V+
\pi \left(J_q^{2}\displaystyle\frac{\tanh\gamma J_{3}^\prime }{\gamma}
-\displaystyle\frac{\gamma J_{3}^\prime
-\tanh\gamma J_{3}^\prime}{\gamma^{2}\sinh\gamma}\right)\right)
\right\vert^{\frac{\sinh\gamma J_{3}^\prime}
{\sqrt{\gamma\sinh\gamma}}=+J_q}
_{\frac{\sinh\gamma J_{3}^\prime}{\sqrt{\gamma\sinh\gamma}}=-J_q}~,
\end{equation}
where V is the volume of the manifold $M_0^q$,
\begin{equation}
\begin{array}{rcl}
V&=&\displaystyle\int_{M_0^q}dJ_{1}^\prime dJ^\prime_{2}dJ^\prime_{3}\\[4mm]
&=&\left.\left(\pi J_q^{2}J^\prime_{3}-\displaystyle\frac{\pi}{2}
\displaystyle\frac{\sinh 2\gamma J^\prime_{3}-
2\gamma J^\prime_{3}}{2\gamma^{2}\sinh\gamma}\right)
\right\vert^{\frac{\sinh\gamma
J^\prime_{3}}{\sqrt{\gamma\sinh\gamma}}=+J_q}
_{\frac{\sinh\gamma J^\prime_{3}}{\sqrt{\gamma\sinh\gamma}}=-J_q}~.
\end{array}
\end{equation}
Then, we have
\begin{equation}\label{5.2.37}
\int_{M_0^q}\Omega_q=-4\pi\displaystyle\frac{\sinh^{-1}
(\sqrt{\gamma\sinh\gamma}J_q)}{\gamma}~.
\end{equation}
Setting
\begin{equation}
\frac{\sinh^{-1}(\sqrt{\gamma\sinh\gamma}J_q)}{\gamma}=J~,
\end{equation}
we get
\begin{equation}
-(2\pi)^{-1}\int_{M_0^q}\Omega_q=4\pi (2\pi)^{-1}J=2J~,
\end{equation}
which must be an integer if $\{-(2\pi)^{-1}\Omega_q\}$ is integrable.
Therefore, $2J\in Z$ and J is an integer or half integer.
It is clear now that $J_q$ takes some special values according to J,
\begin{equation}\label{5.e274}
J_q=\displaystyle\frac{\sinh\gamma J}{\sqrt{\gamma\sinh\gamma}}~.
\end{equation}
Comparing Eq.(\ref{5.e267}) with Eq.(\ref{5.e274}), we know that here b
should be integer or half integer.
For a suitable polarization let us consider the linear
frame fields $ X_{z_{\pm}}$,
\begin{equation}
X_{z_{\pm}}=- i Q_{\pm}^{-1}\displaystyle\frac{\partial}{\partial
\overline{z}_{\pm}}~.
\end{equation}
For each $x \in U_{+}\cap U_{-}$, we have
$$ X_{z_{-}}=-z_{+}^{-2}X_{z_{+}}~.$$
Thus, $X_{z_+}$ and $X_{z_-}$ span a complex dis\-tri\-bu\-tion $F$
on $M_0^q$ and $F$ is a polari\-za\-tion of symplectic manifold
\hspace{0.5mm} ($M_0^q$, $\Omega_q$).\hspace{0.5mm} Further,
\begin{equation}\label{5.2.42}
i\Omega_q\left(X_{z_\pm},\overline{X}_{z_\pm}\right)=-\frac{1}{2J_q}
\sqrt{J_q^{2}\gamma^{2}\frac{(1-z_\pm\overline{z}_\pm)^2}
{(1+z_\pm\overline{z}_\pm)^2}+\displaystyle\frac{\gamma}{\sinh \gamma}}
~(1+z_\pm\overline{z}_\pm)^2 >0~,
\end{equation}
implying that $F$ is a complete strongly admissible positive polarization of
($M_0^q$, $\Omega_q$).
To get the quantum operator expressions for $J_{i}^\prime$ ($i=1,~2,~3$),
we start with the quantum operators of $p$ and $z$. For the polarization
preserving functions $p$ and $z$ Eq.(\ref{5.e268}) yields
\begin{equation}
[X_{p_\pm},X_{z_\pm}]=0~ ,
\end{equation}
whose quantum counterparts are
\begin{equation}
\hat{p}_\pm=-\frac{\partial}{\partial z_\pm}~,~~~
\hat{z}_\pm=z_\pm~,
\end{equation}
where the terms with derivative $\displaystyle\frac{\partial}{\partial
\overline{z}}$ have been omitted as the section space is covariantly constant
along the
polarization F, i.e., the quantum representation space
is the holomorphic section space. Thus, the quantum commutator of
$\hat{p}$ and $\hat{z}$ is
\begin{equation}\label{5.e278}
[\hat{z}_\pm,\hat{p}_\pm]=1~.
\end{equation}
Making use of Eqs.(\ref{5.e221}) and (\ref{5.e278}), we obtain the quantum
operators with suitable
ordering
\begin{equation}\label{5.e280}
\begin{array}{rcl}
\hat{J}_{1}^\prime&=&-\displaystyle\frac{1}{\sqrt{\gamma\sinh\gamma}}
\left(\cosh\left(\frac{\gamma}{2}z_\pm\frac{\partial}{\partial z_\pm}
\right)z_\pm\sinh\left(\frac{\gamma}{2}\left(-z_\pm\frac{\partial}
{\partial z_\pm}+2J\right)\right)\right.\\[4mm]
&&~~~~~~\displaystyle+\cosh\left.\left(\frac{\gamma}{2}\left(-z_\pm
\frac{\partial}{\partial z_\pm}+2J\right)\right)\frac{1}{z_\pm}
\sinh\left(\frac{\gamma}{2}z_\pm\frac{\partial}{\partial z_\pm}\right)
\right)~,\\[4mm]
\hat{J}_{2}^\prime&=&\displaystyle\frac{\mp i}{\sqrt{\gamma\sinh\gamma}}
\left(\cosh\left(\frac{\gamma}{2}z_\pm\frac{\partial}{\partial z_\pm}
\right)z_\pm\sinh\left(\frac{\gamma}{2}\left(-z_\pm\frac{\partial}
{\partial z_\pm}+2J\right)\right)\right.\\[4mm]
&&~~~~~~\displaystyle\left.-\cosh\left(\frac{\gamma}{2}\left(-z_\pm\frac{\partial}
{\partial z_\pm}+2J\right)\right)\frac{1}{z_\pm}\sinh\left(
\frac{\gamma}{2}z_\pm\frac{\partial}{\partial
z_\pm}\right)\right)~,\\[4mm]
\hat{J}_{3}^\prime&=&\displaystyle\mp\left(-z_\pm\frac{\partial}{\partial
z_\pm}+J\right)~. \end{array}
\end{equation}
Equation (\ref{5.e280}) yields a realization of the quantum group $SU_q(2)$
\begin{equation}
\begin{array}{rcl}
[\hat{J}_1^\prime,\hat{J}_2^\prime]&=&-\displaystyle\frac{i\sinh\gamma}{2\gamma}
[2\hat{J}_3^\prime]_q~,\\[4mm]
[\hat{J}_2^\prime,\hat{J}_3^\prime]&=&-i \hat{J}^\prime_1~,~~~~
[\hat{J}^\prime_3,\hat{J}^\prime_1]=-i\hat{J}^\prime_2~;
\end{array}
\end{equation}
or
\begin{equation}\label{5.2.48}
[\hat{J}_+^\prime,\hat{J}^\prime_{-}]=-\displaystyle\frac{\sinh(\gamma)}{\gamma}
[2\hat{ J}^\prime_{3}]_q~,~~~
[\hat{J}^\prime_{3},\hat{J}^\prime_\pm]=\mp\hat{ J}^\prime_\pm~.
\end{equation}
The Hopf algebra structure is of the form
\begin{equation}
\begin{array}{l}
\Delta(\hat{J}_3')=\hat{J}_3'\otimes 1+1\otimes \hat{J}_3' ~,\\
\Delta(\hat{J}_\pm')=\hat{J}_\pm'\otimes e^{-\gamma\hat{J}_3'}
+e^{\gamma\hat{J}_3'}
\otimes \hat{J}_3' ~,\\
\epsilon(\hat{J}_\pm') =\epsilon(\hat{J}_3') =0~,\\
S(\hat{J}_\pm')=-q^{\pm 1}\hat{J}_\pm'~,~~~~S(\hat{J}_3')=-
\hat{J}_3' ~.
\end{array}
\end{equation}
For the quantum symmetric top system, making use of the quantum group
$SU_{q}(2)$ symmetry, we write the Heisenberg equation as
\begin{equation}\label{5.2.50}
\begin{array}{rcl}
i\hbar\dot{\hat{J}^\prime}_1&=&[\hat{J}^\prime_1,\hat{H_q}]=
\displaystyle\frac{i\hbar^2(I-I_3)}{2II_3}
\left(\hat{J}^\prime_2\hat{J}^\prime_3+\hat{J}^\prime_3
\hat{J}_2^\prime\right)~,\\[3mm]
i\hbar\dot{\hat{J}^\prime}_2&=&[\hat{J}^\prime_2,\hat{H_q}]=
\displaystyle\frac{i\hbar^2(I_3-I)}{2II_3}
\left(\hat{J}^\prime_1\hat{J}^\prime_3+\hat{J}^\prime_3
\hat{J}^\prime_1\right)~,\\[3mm]
i\hbar\dot{\hat{J}^\prime}_3&=&[\hat{J}^\prime_3,\hat{H_q}]=0~,
\end{array}
\end{equation}
or
\begin{equation}\label{5.2.50.2}
\begin{array}{rcl}
\dot{\hat{J}^\prime}_1&=&\displaystyle\frac{\hbar(I-I_3)}{2II_3}
\left(\hat{J}^\prime_2\hat{J}^\prime_3+\hat{J}^\prime_3
\hat{J}^\prime_2\right)~,\\[3mm]
\dot{\hat{J}^\prime}_2&=&\displaystyle\frac{\hbar(I_3-I)}{2II_3}
\left(\hat{J}^\prime_1\hat{J}_3^\prime+\hat{J}^\prime_3
\hat{J}^\prime_1\right)~,\\[3mm]
\dot{\hat{J}^\prime}_3&=&0~,
\end{array}
\end{equation}
which are the same as those for the quantum symmetric top system with
standard Hamiltonian. Thus, the quantum symmetric top system with
deformed Hamiltonian and the quantum symmetric top system with
standard Hamiltonian obey the same Heisenberg
equation in quantum mechanics.
The quantum counterparts \cite{Zachos} of Eq.(\ref{5.F1}) are
\begin{equation}\label{5.F3}
\begin{array}{rcl}
\hat{J}_+^\prime&=&\displaystyle\frac{1}{\sqrt{\gamma\sinh\gamma}}
\sqrt{\displaystyle\frac{\sinh\gamma(\hat{J}_0+\hat{J}_3)
\sinh\gamma(\hat{J}_0-\hat{J}_3+1)}
{(\hat{J}_0+\hat{J}_3)(\hat{J}_0-\hat{J}_3+1)}}
\hat{J}_+~,\\[4mm]
\hat{J}_-^\prime&=&\displaystyle\frac{1}{\sqrt{\gamma\sinh\gamma}}
\hat{J_-}\sqrt{\displaystyle\frac{\sinh\gamma(\hat{J}_0+\hat{J}_3)
\sinh\gamma(\hat{J}_0-\hat{J}_3+1)}
{(\hat{J}_0+\hat{J}_3)(\hat{J}_0-\hat{J}_3+1)}}~,\\[4mm]
\hat{J}_3^\prime&=&\hat{J}_3~.
\end{array}
\end{equation}
It is well-known that the stationary states of the symmetric top can be given
by the Wigner $D$-functions, $D^{J}_{MK}$ \cite{Biedenharn81},
\begin{equation}\label{5.2.17}
D^{J}_{MK}(\alpha,\beta,\gamma)=e^{iM\alpha+iK\gamma}d^J_{MK}(\beta)~,
\end{equation}
where
\begin{equation}\label{5.253}\begin{array}{rcl}
d^J_{MK}(\beta)&=&\left((J+M)!(J-M)!(J+K)!(J-K)!\right)^{1/2}\times\\
&
&\displaystyle\times\sum_\nu\left((-1)^\nu(J-M-\nu)!(J+K-\nu)!(\nu+M-K)!~\nu!~
\right)^{-1}\times\\
& &~~~~~\times\left(\cos\beta/2\right)^{2J+K-M-2\nu}
\left(-\sin\beta/2\right)^{M-K-2\nu}
\end{array}
\end{equation}
($\alpha$, $\beta$ and $\gamma$ are the Euler angles). In terms of the Euler
angles, $\hat{J}_i$'s can be expressed as
\begin{equation}\label{5.F2}
\begin{array}{rcl}
\hat{J_+}&=&e^{i\alpha}\left(i\cot\beta\displaystyle\frac{\partial}
{\partial\alpha}+\frac{\partial}{\partial\beta}-\frac{i}{\sin\beta}
\frac{\partial}{\partial\gamma}\right)~,\\[3mm]
\hat{J_-}&=&e^{-i\alpha}\left(i\cot\beta\displaystyle\frac{\partial}
{\partial\alpha}-\frac{\partial}{\partial\beta}-\frac{i}{\sin\beta}
\frac{\partial}{\partial\gamma}\right)~,\\[3mm]
\hat{J_3}&=&-i\displaystyle\frac{\partial}{\partial\alpha}~,
\end{array}
\end{equation}
resulting in the differential equations
\begin{equation}\label{5.e253}
\begin{array}{rcl}
\hat{J}_\pm D^{J}_{MK}&=&\sqrt{(J\mp M)(J\pm M-1)}D^{J}_{M\pm 1,K}~,\\[3mm]
\hat{J}_3D^{J}_{MK}&=&K D^{J}_{MK}~,
~~~~\hat{J}_zD^{J}_{MK}=M D^{J}_{MK}~,
\end{array}
\end{equation}
where $\hat{J}_z$ is the projection of ${\bf \hat{J}}$ onto the $z$-axis
of lab-fixed coordinate system.
Using Eqs.(\ref{5.F2}) and (\ref{5.F3}), one can verify that
$\hat{J}_i^\prime$'s satisfy
the following differential equations,
\begin{equation}
\begin{array}{rcl}
\hat{J}_\pm^\prime D^{J}_{MK}&=&\sqrt{[J\mp M][J\pm M-1]}
D^{J}_{M\pm 1,K}~,\\[3mm]
\hat{J}_3^\prime D^{J}_{MK}&=&K D^{J}_{MK}~,
~~~~\hat{J}_z^\prime D^{J}_{MK}=M D^{J}_{MK}~.
\end{array}
\end{equation}
{}From the above equations, it is not difficult to verify that the
eigenvalues of the deformed Hamiltonian (\ref{5.ssss}) are
\begin{equation}\label{5.2.52}
E_{JK}^q=\frac{\hbar^2}{2I}[J][J+1]+\frac{I-I_3}{2II_3}\hbar^2K^2~.
\end{equation}
\section{Integrable lattice model}
The integrable lattice model \cite{Baxter82}, defined on a two-dimensional
square
lattice (Fig.1), can be divided into two types: vertex model
\cite{Baxter73a}--\cite{Baxter73c} and
Solid-On-Solid (S.O.S) model \cite{Andrews84}--\cite{Date87}. State variables
of vertex model are located
on the edges. We associate the Boltzmann weight with each vertex
configuration defined by the state variables on the four
edges joining together at the vertex (Fig.2). The degrees
of freedom of S.O.S model are located on the sites and interact
through ``interaction-round-face'' around each plaquette.
The Boltzmann weight is
assigned to each unit face depending on the state variable configuration
round the face (Fig.12).
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,15)(0,7)
\multiput(35,10)(0,8){2}{\line(1,0){8}}
\multiput(35,10)(8,0){2}{\line(0,1){8}}
\put(32,9){$a_i$}
\put(32,17.7){$a_l$}
\put(44,9){$a_j$}
\put(44,17.7){$a_k$}
\end{picture}
\hspace{8em}Fig.12. Boltzmann weight $w(a_i,a_j,a_k,a_l)$
\hspace{11.4em}of the S.O.S model.
A precise
study of the thermodynamics of those models indicates that they undergo a
second phase transition at a certain critical temperature $T=T_c$. This fact,
combined with the short range of interactions, implies that in a suitable
critical continuum limit, the system be locally scale, rotation and
translation invariant \cite{Cardy87}.
Considerable progress has been made recently in understanding the
structure of those models, from their connection with
quantum groups and conformal field theory. Integrability
of those models is ensured by local Boltzmann weights satisfying the YBE.
Previously known models have been generalized in several directions.
In particular, the celebrated
six-vertex model and the related XXZ spin 1/2 quantum chain have been
recognized to be the first hirarches, involving either higher spin
representations of $su(2)$ or higher rank algebras or both. This
progress has been made possible by the algebraic formulation of the Yang-
Baxter integratibility condition in the quantum group form. Concept
of quantum group becomes a major theme of the study \cite{Saleur90,Jimbo90k}.
\subsection{Vertex model}
Let us consider a two-dimensional square lattice, whose state variables
are located
on the edges. We associate the Boltzmann weight with each vertex configuration.
The configuration is defined by the state variables say, $i,~j,~k,~l$ on the
four
edges joining together at the vertex. An example is the
$6$-vertex model, for which we set a definite direction by an arrow on each
edge of the
lattice. Four edges meet at each lattice point, and so there are $12$
distinct types of combinations of arrows (Fig.13). We only consider the
configurations
with equal number of incoming and outcoming arrows.
For type $j$ configuration $(j=1,~2,~\cdots,~6)$, we assign energy
$\epsilon(j)$ which is assumed
to be invariant under simultaneous inversion of the direction
of all arrows, so that
\begin{equation}
\epsilon(1)=\epsilon(2)~,~~~~ \epsilon(3)=\epsilon(4)~,~~~~\epsilon(5)=
\epsilon(6)~.
\end{equation}
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,15)(-3,1)
\multiput(0,10)(13,0){6}{\line(1,0){8}}
\multiput(4,6)(13,0){6}{\line(0,1){8}}
\multiput(0,10)(26,0){3}{\vector(1,0){2}}
\multiput(6,10)(26,0){2}{\vector(1,0){1}}
\put(67,10){\vector(-1,0){1}}
\multiput(15,10)(26,0){3}{\vector(-1,0){1}}
\multiput(21,10)(26,0){2}{\vector(-1,0){2}}
\put(71,10){\vector(1,0){1}}
\put(60,10){\vector(-1,0){2}}
\multiput(4,6)(0,5){2}{\vector(0,1){2}}
\multiput(17,8)(0,5){2}{\vector(0,-1){1}}
\multiput(30,8)(0,5){2}{\vector(0,-1){1}}
\multiput(43,6)(0,5){2}{\vector(0,1){2}}
\put(56,8){\vector(0,-1){1}}
\put(56,12){\vector(0,1){1}}
\put(69,6){\vector(0,1){2}}
\put(69,14){\vector(0,-1){2}}
\put(4,3){$1$}
\put(17,3){$2$}
\put(30,3){$3$}
\put(43,3){$4$}
\put(56,3){$5$}
\put(69,3){$6$}
\end{picture}
\hspace{10em}Fig.13. The six-vertex model.
This completely defines the $6$-vertex model. With each configuration on the
entire lattice,
we associate a total energy $E$ with
\begin{equation}
E=\sum_{i=1}^6 n_i\epsilon(i)~,
\end{equation}
where $n_i$ is the number of type $i$ vertex in the given configuration.
The partition function $Z_N$ and the free energy per site $f$ are then given by
\begin{equation}
\begin{array}{rcl}
& &Z_N=\sum\exp(-\beta E)~,~~~~\beta=1/k_BT~,\\[2mm]
& &f=k_BT\displaystyle\lim_{N\to\infty}N^{-1}\log Z_N~,
\end{array}
\end{equation}
where $k_B$ is the Boltzmann constant, $T$ is the (Kelvin) temperature and $N$
is the
number of lattice site and the summation is taken over all configurations of
arrows.
For a given configuration of lattice, we
consider a horizontal row of the lattice and the adjacent vertex edges.
Let $\alpha=\{\alpha_i,\cdots,\alpha_n\}$ be the state variables on the
lower row of the vertical edges, $\alpha'=\{\alpha'_i,\cdots,\alpha'_n\}$ be
the
state variables on the upper row and $\beta=\{\beta_i,\cdots,\beta_n\}$ be the
state variables on the horizontal edges.
We adopt a Hamiltonian picture, in which ``time'' flows upward
on the lattice, and the various configurations of vertical links are considered
being independent possible states of the system at a given time. Time evolution
is carried out by the row-to-row transfer matrix ${\cal V}^{(n)}$ (Fig.3),
whose
matrix elements are ${\cal V}^{(n)}_{\alpha,\alpha'}$ is defined by
\begin{equation}\label{3.2.11}
{\cal V}^{(n)}_{\alpha,\alpha'}=\sum_{\beta_1\cdots\beta_n}
w(\beta_1,\alpha_1,\beta_2,\alpha_1')
w(\beta_2,\alpha_2,\beta_3,\alpha_2')
\cdots w(\beta_n,\alpha_n,\beta_1,\alpha_n')~,
\end{equation}
where $w(\beta_i,\alpha_i,\beta_i',\alpha_i')$ is Boltzmann weight
of the vertex
$$w(\beta_i, \alpha_i,\beta_i', \alpha_i')=
\exp\left(-\epsilon(\beta_i,\alpha_i,\beta_i',\alpha_i')/k_BT\right)~.$$
For the general case, the degrees of freedom denoted by $\alpha$ and $\beta$
may be of a rather general nature and take their values in a discrete set.
If the vertical ones take $d$ independent values, the space spanned
by those states is $V^{\otimes n}$, where $V$ is the one-body vector space:
$V={\bf C}^d$ for the $d$-state model.
In terms of the transfer matrix ${\cal V}^{(n)}$, the partition function $Z_N$
and the free
energy per site $f$ are given by
\begin{equation}
\begin{array}{rcl}
& &Z_N=\displaystyle\sum_{\alpha^{(1)}\alpha^{(2)}\cdots \alpha^{(m)}}
{\cal V}^{(n)}_{\alpha^{(1)}\alpha^{(2)}} {\cal V}^{(n)}_{\alpha^{(2)}
\alpha^{(3)}}\cdots {\cal V}^{(n)}_{\alpha^{(m-1)}\alpha^{(m)}}
{\cal V}^{(n)}_{\alpha^{(m)}\alpha^{(1)}}
={\rm Tr}\left({\cal V}^{(n)}\right)^m~,\\
& &f=-k_BT\displaystyle\lim_{N\to\infty}N^{-1}\log Z_N~.
\end{array}
\end{equation}
The consistency condition of the transfer matrix ${\cal V}^{(n)}$ yields
the YBE (Fig.14),\\ \\
$\displaystyle\sum_{\gamma\beta''\bar{\beta}''}
w(\beta,\alpha,\beta'',\gamma)
w(\bar{\beta},\gamma,\bar{\beta}'',\alpha')
w(\bar{\beta}'',\beta'',\beta',\bar{\beta}')$
\begin{equation}\label{3.2.18}
=\sum_{\gamma\beta''\bar{\beta}''}
w(\bar{\beta},\beta,\beta'',\bar{\beta}'')
w(\beta'',\alpha,\beta',\gamma)w(\bar{\beta}'',\gamma,\bar{\beta}',\alpha')~.
\end{equation}
\setlength{\unitlength}{4pt}
\thicklines
\begin{picture}(50,32)(-18,0)
\multiput(3,5)(30,0){2}{\line(1,1){21}}
\multiput(23,5)(30,0){2}{\line(-1,1){21}}
\multiput(10,5)(37,0){2}{\line(0,1){21}}
\put(27,15){$=$}
\multiput(3,2)(30,0){2}{$\beta$}
\multiput(10,2)(37,0){2}{$\alpha$}
\multiput(23,2)(30,0){2}{$\beta'$}
\multiput(3,26.7)(30,0){2}{$\bar{\beta}$}
\multiput(10,26.7)(37,0){2}{$\alpha'$}
\multiput(23,26.7)(30,0){2}{$\bar{\beta}'$}
\multiput(7,15)(41,0){2}{$\gamma$}
\multiput(12,19)(30,0){2}{$\bar{\beta}''$}
\multiput(12,10)(30,0){2}{$\beta''$}
\end{picture}
\hspace{8em}Fig.14. YBE for the vertex model.
The quantities
$w(\beta,\alpha,\beta',\alpha')$ can be straightforwardly
interpreted as an operator ${\cal R}$ in the tensor product space
$V^{\otimes 2}$. Thus Eq.(\ref{3.2.18}) can be rewritten in the form
\begin{equation}\label{3.cd1}
{\cal R}_{12}(u){\cal R}_{13}(u+v){\cal R}_{23}(v)={\cal R}_{23}(v)
{\cal R}_{13}(u+v){\cal R}_{12}(u)~.
\end{equation}
Corresponds to the 6-vertex model, the solution of the YBE has the form
\begin{equation}\label{3.I8}
\begin{array}{rcl}
{\cal R}(u) &=&\displaystyle\frac{1}{4}\sum_{i=0}^3w_i\sigma^i
\otimes \sigma^i\\[7mm]
&=&\rho\left[\begin{array}{cccc}
a& & & \\
& b & c & \\
& b & c & \\
& & &a
\end{array}\right]~,
\end{array}
\end{equation}
where
$$\begin{array}{rcl}
& &a=\sin(\eta+u)=\displaystyle\frac{w_0+w_3}{4\rho}=\frac{1}{\rho}
\exp(-i\beta\epsilon(1))~,\\[3mm]
& &b=\sin
u=\displaystyle\frac{w_0-w_3}{4\rho}=\frac{1}{\rho}\exp(-i\beta\epsilon(3))~,\\[3mm]
&
&c=\sin\eta=\displaystyle\frac{w_1}{4\rho}=\frac{1}{\rho}\exp(-i\beta\epsilon(5))~,
\end{array}$$
here $\sigma^i~(i=1,~2,~3)$ are the Pauli matrices, $\sigma^0={\bf 1}$,
$w_1=w_2$,
$\eta$ is a free parameter and $\rho$ is an overall irrelevant factor.\\
Notice that for $u = 0$,
\begin{equation}
{\cal R}(0) = \rho \sin \eta\left[\begin{array}{cccc}
1 & 0 & 0 & 0\\
0 & 0 & 1 & 0\\
0 & 1 & 0 & 0\\
0 & 0 & 0 & 1 \end{array}
\right]= \rho \sin \eta\cdot P~,
\end{equation}
where $P$ is the transposition operator in $V\otimes V$,
\begin{equation}
\begin{array}{rcl}
P:~~V_1\otimes V_2&\to& V_2\otimes V_1~,\\
P:~~a\otimes b&\to& b\otimes a~,
\end{array}
\end{equation}
and that for $u$ and $\eta$ small, $R\approx u${\bf 1}$ + \eta P$.
The introduction of the spectral parameter $u$ in Eq.(\ref{3.I8}) makes the
transfer
matrix ${\cal V}^{(n)}$ $u$ dependent.
The spectral parameter
dependent transfer matrix ${\cal V}^{(n)}(u)$ may be regarded as the generating
function of conserved quantities. The space $V^{\otimes m}$ on which
${\cal V}^{(n)}(u)$ acts, is viewed as the Hilbert space of a quantum
one-dimensional
system. Upon introducing the operators
\begin{equation}
{\cal H}_i = \frac{\partial^i}{\partial u^i} \log {\cal V}^{(n)}(u)|_{u=0}~,
\end{equation}
the commutation of ${\cal V}^{(n)}(u)$ and ${\cal V}^{(n)}(v)$ implies the
commutation of the infinite set of ${\cal H}_{i}$
\begin{equation}
\left[{\cal H}_i, {\cal H}_j\right] =0~.
\end{equation}
In particular, if ${\cal H}_{1}$ is regarded as the Hamiltonian
of the quantum system, there is an infinite number of conserved
quantities, commuting with ${\cal H}_1$.
Let $d\times d$ matrices $t_{\beta\beta'}$ be
\begin{equation}
(t_{\beta\beta'})_{\alpha'\alpha}=w(\beta,\alpha,\beta',\alpha')~.
\end{equation}
In terms of the matrices $t_{\beta\beta'}$, for the six-vertex model
\begin{equation}
\frac{d}{du}{\cal V}^{(n)}(u)|_{u=0}=\sum_{i=1}^nt_{\beta_1\beta_2}(0)\otimes
t_{\beta_2\beta_3}(0)\otimes\cdots\otimes \dot{t}_{\beta_i\beta_{i+1}}(0)
\otimes\cdots\otimes t_{\beta_n\beta_{1}}(0)~.
\end{equation}
Thus, the Hamiltonian of the $6$-vertex model is
\begin{equation}
{\cal H}_{1} =\left({\cal V}^{(n)}(0)\right)^{-1}\dot{\cal V}^{(n)}(0)=
\sum^{n}_{i=0} 1 \otimes \cdots
\otimes h_{i,i+1} \otimes \cdots \otimes 1~,
\end{equation}
where $h_{i,i+1}$ acts on the $i$ and $i+1$-th vertical
variables, i.e., in the space $V_i \bigotimes V_{i+1}$
\begin{equation}
\begin{array}{rcl}
h_{i,i+1} &=& \displaystyle\frac{1}{\rho\sin\eta}P\dot{\cal
R}(0)=\frac{1}{\sin\eta}
\left[\begin{array}{cccc}
\cos \eta & & & \\
& 0 & 1 & \\
& 1 & 0 & \\
& & & \cos \eta \end{array}
\right]\\[8mm]
& =&\displaystyle \frac{1}{2 \sin \eta}\left(\sigma^1 \otimes \sigma^1 +
\sigma^2 \otimes\sigma^2+(1 \otimes 1 + \sigma^3 \otimes
\sigma^3)\cos\eta\right)~.
\end{array}
\end{equation}
Thus, up to an irrelevant constant,
\begin{equation}\label{3.0038}
{\cal H}_1 = \frac{1}{2 \sin \eta} \sum^{n}_{i=1}\left(\sigma^1_i\otimes
\sigma^1_{i+1} + \sigma^2_i \otimes \sigma^2_{i+1} +\cos\eta
\sigma^3_i \otimes\sigma^3_{i+1}\right)~,
\end{equation}
where we have used the notation
$$\sigma^i_k=1_{(1)}\otimes 1_{(2)}\otimes\cdots\otimes 1_{(k-1)}\otimes
\sigma^i_{(k)}\otimes 1_{(k+1)}\otimes\cdots\otimes 1_{(n)}~.$$
This is the Hamiltonian of a (periodic) chain of $\displaystyle
\frac{1}{2}$-spin,
$s_i=\displaystyle\frac{1}{2}\sigma_i$ interacting with an anisotropic``XXZ"-
interaction. Thus, we see that the transfer matrix ${\cal V}^{(n)}(u)$ of
the $6$-vertex model commutes with the XXZ
spin-$\displaystyle\frac{1}{2}$ Hamiltonian.
Alternatively we may say that the latter is generated by the former in a
``very anisotropic limit'' where the lattice spacing in the vertical direction
is
let to zero as well as the spectral parameter,
\begin{equation}
{\cal V}^{(n)}(u)=1+u{\cal H}_1~.
\end{equation}
\setlength{\unitlength}{4pt}
\thicklines
\begin{picture}(50,40)(-20,15)
\multiput(25,20)(-5,5){2}{\line(1,1){25}}
\multiput(25,20)(25,25){2}{\line(-1,1){5}}
\multiput(35,20)(-15,15){2}{\line(1,1){15}}
\multiput(35,20)(15,15){2}{\line(-1,1){15}}
\multiput(45,20)(5,5){2}{\line(-1,1){25}}
\multiput(45,20)(-25,25){2}{\line(1,1){5}}
\multiput(25,50)(10,0){3}{\line(1,1){3}}
\multiput(25,50)(10,0){3}{\line(-1,1){3}}
\put(5,25){time}
\put(9,28){\vector(0,1){15}}
\end{picture}
\hspace{7.5em}Fig.15. Diagonal-to-diagonal transfer matrix
\hspace{10.7em}for the vertex model.
Now we consider the six-vertex model, but with a transfer matrix
propagating in the diagonal direction (Fig.15). We restrict ourselves solely
to free boundary condition (depending on the parity of $n$).
The diagonal-to-diagonal transfer matrix ${\cal V}^{(n)}(u)$ acts on
${\cal H} =(V(1))^{\otimes n}$ and related with a gauge transformed $R$
matrix. It is well-known that there are some symmetries of the $R$ matrix.
The symmetries enable us to modify
the $R_{1221}$ and $R_{2112}$ entries of the matrix, retaining their
product unchanged, and preserving the YBE \cite{Akutsu87}. This is
indeed true, and
\begin{equation}\label{3.I11}
{\cal R}(u) = \left[\begin{array}{cccc}
a & & & \\
& b & ce^{xu} & \\
& ce^{-xu} & b & \\
& & & a
\end{array}\right]
\end{equation}
also satisfies the YBE. The merit of this form is that it
leads to a non trivial limit as $u \rightarrow \pm i\infty$, for
$x = -i$,
\begin{equation}\label{3.666}
{\cal R}(-i\infty) \sim \left[\begin{array}{cccc}
1 & & &\\
& q^{-1} & 0 & \\
& 1 - q^{-2} & q^{-1}& \\
& & & 1 \end{array}
\right]~,~~~~
{\cal R}(+i \infty) \sim \left[\begin{array}{cccc}
1 & & & \\
& q & 1-q^{2} & \\
& 0 & q & \\
& & & 1 \end{array}
\right]~,
\end{equation}
where $q = e^{i\eta}$. There are other more general type of such
"gauge transformations" of the $R$-matrix. Corresponding to the generalized
form
of $R$ matrix (\ref{3.I11}), the six-vertex model is generalized nicely
to a $d$-state model. With the gauge transformed $R$ matrix
\begin{equation}
{\cal R}(u) = \left[\begin{array}{cccc}
\sin(\eta+u) & & & \\
& \sin u & \sin\eta ~e^{iu} & \\
& \sin\eta~ e^{-iu} & \sin u & \\
& & &\sin(\eta+u)
\end{array}\right]~,
\end{equation}
the Boltzmann weights read
\begin{equation}
a=\sin(\eta+u)~,~~~~b=\sin u~,~~~~c_1=\sin\eta e^{iu}~,~~~~c_2=\sin\eta
e^{-iu}~.
\end{equation}
In terms of the above gauge transformed weights the diagonal-to-diagonal
transfer matrix is
\begin{equation}
{\cal V}^{(n)}(u)=\prod_{i=0}^{\{\frac{n-1}{2}\}}\left(
\sin(\eta+u)1-\sin u~ e_{2i+1}\right)
\prod_{i=1}^{\{\frac{n}{2}\}}
\left(\sin(\eta+u)1-\sin u~ e_{2i}\right)~,
\end{equation}
where
\begin{equation}\label{3.I212}
\begin{array}{rcl}
e_i&=&1_{(1)}\otimes 1_{(2)}\otimes\cdots\otimes 1_{(i-1)}\otimes e_{(i,i+1)}
\otimes 1_{(i+2)}\otimes\cdots\otimes 1_{(n)}~,\\[7mm]
e&=&\left[\begin{array}{cccc}
0 &0 &0 &0\\
0 &q^{-1}&-1 &0\\
0 &-1 &q &0\\
0 &0 &0 &0\end{array}\right]\\[6mm]
&=&-\displaystyle\frac{1}{2}\left(\sigma^1\otimes\sigma^1+\sigma^2\otimes
\sigma^2+\cos\eta\sigma^3\otimes\sigma^3\right)+\frac{1}{2}\cos\eta 1-
\frac{i}{2}\sin\eta \left(\sigma^3\otimes 1-1\otimes\sigma^3\right)~.
\end{array}
\end{equation}
The $e_i$ given in Eq.(\ref{3.I212}) is the quantum analogue of $1 -
P_{i(i+1)}$,
where $P_{i(i+1)}$ is the transposition of the $i$-th and $(i+1)$-th spaces.
It is easy to verify that $e_i$ satisfy the Temperley-Lieb algebra $A_n$
\cite{Lieb},
\begin{equation}
\begin{array}{rcl}
& &e^2_i=\left[2 \right] e_i~,\\
& &e_i e_{i+1} e_i=e_i~,\\
& &e_i e_j=e_j e_i~, ~~~~{\rm if}~ |i - j| \geq 2~.
\end{array}
\end{equation}
The very anisotropic limit of the diagonal-to-diagonal transfer matrix yields
\begin{equation}
{\cal V}^{(n)}(u)=1-\frac{u}{\sin\eta}\sum_{i=1}^{n}e_i~.
\end{equation}
Thus, we obtain the Hamiltonian of the XXZ spin-1/2 chain as
\begin{equation}\label{3.0039}
\begin{array}{rcl}
{\cal H}_1
&=&-\displaystyle\frac{1}{\sin\eta}\displaystyle\sum_{i=1}^{n}e_i\\[3mm]
&=&\displaystyle\frac{1}{2\sin\eta}\sum^n_{i=1}\left(\sigma^1_i
\otimes\sigma^1_{i+1} + \sigma^2_i \otimes \sigma^2_{i+1}+
\cos\eta\sigma^3_i \otimes\sigma^3_{i+1}+i\sin\eta(\sigma^3_i
-\sigma^3_{i+1})-\cos\eta\right)~.
\end{array}
\end{equation}
This Hamiltonian is different from the standard XXZ Hamiltonian (\ref{3.0038})
by
the boundary terms.
The quantum group $SU_q(2)$ is known as
\begin{equation}
\begin{array}{l}
[S^3,S^\pm]=\pm S^\pm~,\\[1mm]
[S^+,S^-]=[2S^3]~,\\[1mm]
\Delta(S^3)=S^3\otimes 1+1\otimes S^3~,\\[1mm]
\Delta(S^\pm)=S^\pm\otimes q^{-S^3}+q^{S^3}\otimes S^\pm~,\\
\epsilon(S^3)=0=\epsilon(S^\pm)~,\\
S(S^3)=-S^3~,~~~~S(S^\pm)=-q^{\pm 1}S^\pm~.
\end{array}
\end{equation}
For $j=\displaystyle\frac{1}{2}$ the generators of $SU_q(2)$ coincide with
their $q=1$ limit, i.e., the $s^i=\displaystyle\frac{1}{2}\sigma^i$ matrices.
Thus, in $\left(V(1)\right)^{\otimes 2}$ we have
\begin{equation}\label{3.a0}
\begin{array}{l}
S^3=s^3_{(1)}\otimes 1_{(2)}+1_{(1)}\otimes s^3_{(2)}~,\\
S^\pm=s^\pm_{(1)}\otimes q^{-s^3_{(2)}}+q^{s^3_{(1)}}\otimes s^\pm_{(2)}~.
\end{array}
\end{equation}
As well known, there is another comultiplication $\Delta'$. In
$\left(V(1)\right)^{\otimes 2}$, $\Delta'$ is
\begin{equation}\label{3.a1}
\Delta'=P\circ\Delta\circ P~.
\end{equation}
$\Delta'$ and $\Delta$ are related by the universal $R$ matrix of $SU_q(2)$
\begin{equation}\label{3.a2}
\Delta'{\cal R}={\cal R}\Delta~.
\end{equation}
It should be noticed that the explicit form of the universal $R$ matrix in
$\left(V(1)\right)^{\otimes 2}$
\begin{equation}
{\cal R}^{\frac{1}{2}\frac{1}{2}}=q^{\frac{1}{2}}
\left[\begin{array}{cccc}
1 & & &\\
& q^{-1} & 0 & \\
& 1 - q^{-2} & q^{-1}& \\
& & & 1 \end{array}
\right]
\end{equation}
is the same as the first matrix in Eq.(\ref{3.666}), i.e.,
\begin{equation}
\lim_{u\to -i\infty}{\cal R}(u)={\cal R}^{\frac{1}{2}\frac{1}{2}}~.
\end{equation}
{}From Eqs.(\ref{3.a0})---(\ref{3.a2}), we come to know that the operator
$\breve{\cal R}=P\circ {\cal R}$ commutes with the generators of the quantum
group $SU_q(2)$
\begin{equation}
[\breve{\cal R},SU_q(2)]=0~.
\end{equation}
In $\left(V(1)\right)^{\otimes 2}$, $\breve{\cal R}$ is
\begin{equation}
\breve{\cal R}^{\frac{1}{2}\frac{1}{2}}=q^{-\frac{1}{2}}\left(q-e\right)~,
\end{equation}
and the operator $e$ commutes with the generators of the quantum group
$SU_q(2)$.
Define the generators of the quantum group $SU_q(2)$ in $\left(V(1)\right)
^{\otimes n}$ as
\begin{equation}
\begin{array}{rcl}
& &S^\pm=\bigtriangleup^{(n)} \left(s^\pm \right)=\displaystyle\sum^{n}_{i=1}
q^{s^3_{(1)}}\otimes q^{s^3_{(2)}}\otimes \cdots \otimes q^{s^3_{(i-1)}}
\otimes s^\pm_{(i)}\otimes q^{-s^3_{(i+1)}}\otimes\cdots \otimes
q^{-s^3_{(n)}}~,\\ [3mm]
& &S^3=\bigtriangleup^{(n)} \left(s^3\right)=\displaystyle\sum^{n}_{i=1}
1_{(1)} \otimes 1_{(2)}\otimes \cdots \otimes 1_{(i-1)} \otimes s^3_{(i)}
\otimes\cdots \otimes 1_{(n)}~,
\end{array}
\end{equation}
and $s^\pm_{(i)},~s_{(i)}$ are located in the $i$-th position, acting on the
$i$-th spin space. Thus, the Hamiltonian (\ref{3.0039}) is obviously
$SU_q(2)$-invariant, i.e.,
\begin{equation}\label{3.I415}
[{\cal H}_1,SU_q(2)]=0~,
\end{equation}
and the vertex model has an $SU_q(2)$ symmetry besides the
obvious $U(1)$ symmetry due to spin conservation. It is important
to note that Eq.(\ref{3.I415}) would not hold for the standard weights
(\ref{3.I8}).
The effect of the gauge transformation can be put in boundary
terms only, but these are crucial as far as symmetries and critical
properties are concerned.
\subsection{S.O.S model}
Another family of integrable model is the S.O.S model. The degrees
of freedom of S.O.S model are located on the sites of a square lattice and
interact
through ``interaction-round-face'' around each plaquette. The Boltzmann weight
is
assigned to each unit face depending on the state variable configuration round
the
face (Fig.12). We denote energy of a face by the state variable configuration
$(a_i,a_j,a_k,a_l)$ as $\epsilon(a_i,a_j,a_k,a_l)$.
The S.O.S model is very general: most of exactly solvable models
can be expressed in the form of S.O.S model. Total energy of the entire lattice
is
\begin{equation}
E=\sum_{\rm all~faces}\epsilon(a_i,a_j,a_k,a_l)~.
\end{equation}
In terms of the Boltzmann weights $w(a_i,a_j,a_k,a_l)$,
the partition function $Z_N$ and the free energy per site $f$ are given by
\begin{equation}
\begin{array}{rcl}
& &Z_N=\displaystyle\sum_{a_1}\cdots\sum_{a_N}\sum_{(i,j,k,l)}
w(a_i,a_j,a_k,a_l)~,\\[3mm]
& &f=k_BT\displaystyle\lim_{N\to\infty}N^{-1}\log Z_N~.
\end{array}
\end{equation}
Notice that the Boltzmann weight are defined up to a gauge transformation that
has no effect on the partition function in the thermodynamic limit
\begin{equation}
w(a_i,a_j,a_k,a_l)\longrightarrow\frac{f(a_i,a_j)g(a_l,a_i)}{f(a_k,a_l)
g(a_j,a_k)}w(a_i,a_j,a_k,a_l)~.
\end{equation}
The constraint on the weights expresses that no strongly fluctuating
configuration is allowed. More precisely, if the $l$'s take integers
values (on a finite or infinite range), we have a nonzero Boltzmann weight iff
any
two neighboring heights around the face differ by $\pm 1$.
The row-to-row transfer matrix ${\cal V}^{(n)}$ for the S.O.S model has matrix
elements (Fig.16)
\begin{equation}
{\cal V}^{(n)}_{a,a'}=\prod_{i=1}^nw(a_i,a_{i+1},a_{i+1}',a_i')~,
\end{equation}
where $a=\{a_1,~a_2,~\cdots,~a_n\},~
a'=\{a'_1,~a'_2,~\cdots,~a'_n\},~a_{n+1}=a_1$,
and $a'_{n+1}=a'_1$.
It generates the time evolution of the system.
Very similar to the vertex model, we get the
YBE as the consistency conditions of the row-to-row transfer matrix
${\cal V}^{(n)}$ (Fig.17),\\ \\
$\displaystyle\sum_{b''}w(a_i,a_{i+1},b'',a_i''|u)w(a_i'',b'',a_{i+1}',a_i'|u+v)
w(b'',a_{i+1},a_{i+1}'',a_{i+1}'|v) $
\begin{equation}\label{3.abc1}
=\sum_{b''}w(a_i'',a_i,b'',a_i'|v)w(a_i,a_{i+1},a_{i+1}'',b''|u+v)
w(b'',a_{i+1}'',a_{i+1}',a_i'|u)~.
\end{equation}
\setlength{\unitlength}{7pt}
\thicklines
\begin{picture}(50,12)(0,7)
\multiput(10,10)(5,0){8}{\line(0,1){5}}
\multiput(10,10)(0,5){2}{\line(1,0){35}}
\put(10,8.5){$a_1$}
\put(15,8.5){$a_2$}
\put(20,8.5){$a_3$}
\put(10,15.9){$a'_1$}
\put(15,15.9){$a'_2$}
\put(20,15.9){$a'_3$}
\put(40,8.5){$a_n$}
\put(45,8.5){$a_{n+1}=a_1$}
\put(40,15.9){$a'_n$}
\put(45,15.9){$a'_{n+1}=a'_1$}
\end{picture}
\hspace{4em}Fig.16. Row-to-row transfer matrix for the S.O.S model.
\setlength{\unitlength}{7pt}
\thicklines
\begin{picture}(50,21)(0,5)
\multiput(10,10)(0,10){2}{\line(1,0){10}}
\multiput(10,10)(10,0){2}{\line(-1,2){2.5}}
\multiput(7.5,15)(10,0){2}{\line(1,2){2.5}}
\put(7.5,15){\line(1,0){10}}
\put(20,10){\line(1,2){2.5}}
\put(22.5,15){\line(-1,2){2.5}}
\put(27,14.5){$=$}
\multiput(32.5,15)(2.5,5){2}{\line(1,-2){2.5}}
\multiput(32.5,15)(2.5,-5){2}{\line(1,2){2.5}}
\multiput(35,10)(0,10){2}{\line(1,0){10}}
\put(37.5,15){\line(1,0){10}}
\put(45,10){\line(1,2){2.5}}
\put(47.5,15){\line(-1,2){2.5}}
\multiput(17.2,14.7)(20,0){2}{$\bullet$}
\put(11.5,17){$u+v$}
\put(14,11.5){$u$}
\put(19.5,14.5){$v$}
\put(34.5,14.5){$v$}
\put(39.3,16.5){$u$}
\put(39,12){$u+v$}
\multiput(5,14.5)(25,0){2}{$a_i''$}
\multiput(23,14.5)(25,0){2}{$a_{i+1}''$}
\multiput(9.5,8.5)(25,0){2}{$a_{i}$}
\multiput(19.5,8.5)(25,0){2}{$a_{i+1}$}
\multiput(9.5,20.5)(25,0){2}{$a'_{i}$}
\multiput(19.5,20.5)(25,0){2}{$a'_{i+1}$}
\put(16,13.5){$b''$}
\put(37.5,13.5){$b''$}
\end{picture}
\hspace{6em}Fig.17. YBE for the the S.O.S model.
Defining the operators $X_i(u)$ for the S.O.S model by
\begin{equation}
X_i(u)_{a,a'}=\prod_{j\not=i}\delta_{a_ja'_j}
w(a_{i-1},a_i,a_{i+1},a_i'|u)~,
\end{equation}
and making use of Eq.(\ref{3.abc1}) yields
\begin{equation}\label{3.a3}
X_i(u)X_{i+1}(u+v)X_i(v)=X_{i+1}(v)X_{i}(u+v)X_{i+1}(u)~.
\end{equation}
The role played by the operator $X_i(u)$ is to evaluate a configuration line by
changing
only one height at site $i$. The solution of the YBE (\ref{3.a3}) is
\begin{equation}
X_i(u)=\sin(\eta+u)1-\sin u ~e_i~.
\end{equation}
In terms of the operators $X_i(u)$, the diagonal-to-diagonal transfer matrix
${\cal V}^{(n)}(u)$ of the S.O.S model (Fig.18) can be written as
\begin{equation}
{\cal V}^{(n)}(u)=\prod_{i=0}^{\{\frac{n-1}{2}\}}X_{2i+1}
\prod_{i=1}^{\{\frac{n}{2}\}}X_{2i}~.
\end{equation}
\setlength{\unitlength}{4pt}
\thicklines
\begin{picture}(50,19)(0,4.5)
\multiput(10,10)(10,0){8}{\line(1,1){5}}
\multiput(15,15)(10,0){8}{\line(1,-1){5}}
\multiput(45,15)(1,-1){6}{\line(1,1){5}}
\multiput(45,15)(1,1){6}{\line(1,-1){5}}
\put(9.5,8.5){$a_1$}
\put(14.5,15.5){$a_2$}
\put(19.5,8.5){$a_3$}
\put(49.5,8.5){$a_i$}
\put(89.5,8.5){$a_n$}
\put(40,15.5){$a_{i-1}$}
\put(49.5,20.5){$a_i'$}
\put(55.5,15.5){$a_{i+1}$}
\end{picture}
\hspace{3em}Fig.18. Diagonal-to-diagonal transfer matrix for the S.O.S model.
Then, in the very anisotropic limit, we obtain the Hamiltonian of the XXZ
model
\begin{equation}\label{3.0040}
{\cal H}_1=-\frac{1}{\sin\eta}\sum_{i=1}^ne_i~.
\end{equation}
Therefore, the S.O.S model also possesses an $SU_q(2)$ quantum symmetry as the
vertex model does.
\subsection{Configuration space}
In the previous sub-sections, we have shown that both the vertex model and the
S.O.S model
are mapped onto spin-$1/2$ quantum XXZ chains. The generators of the quantum
group
$SU_q(2)$, $S^\pm$ and $S^3$, act on the diagonal-to-diagonal lines of a
lattice,
interpreted as a direct product of spin-$1/2$ representations of
quantum group.
According to the Clebsch-Gordan rule for generic $q$, the configuration space
${\cal H}=(V(1))^{\otimes n}$ can be split into a direct sum of
irreducible highest weight representations $V(2j)$, which can be labeled by the
value of the Casimir operator $C$,
\begin{equation}
{\cal H}=\oplus_jw_j\otimes V(2j)~,
\end{equation}
where $w_j$ is a multiplicity space of dimension
$$\Gamma_j^{(n)}=\left(\begin{array}{c}
n\\
\displaystyle\frac{1}{2}n-j\end{array}\right)-
\left(\begin{array}{c}
n\\
\displaystyle\frac{1}{2}n-j-1\end{array}\right)~.$$
Eigenvectors of the diagonal-to-diagonal transfer matrix ${\cal V}^{(n)}$ fill
in representations $V(2j)$ of quantum group $SU_q(2)$, and we denote their
eigenvalues by
$\lambda_j^{(\alpha)},~\alpha=1,~2,~\cdots,~\Gamma_j^{(n)}$. The $SU_q(2)$
symmetry manifests itself through the degeneracies of these
eigenvalues of order $2j+1$. Since $\left({\cal V}^{(n)}\right)^\dagger$ is
equivalent to ${\cal V}^{(n)}$ after spin relabelling, eigenvalues are
real or complex conjugate by pairs.
For $q$ being a root of unity ($q^p=\pm 1$), $\left(V(1)\right)^{\otimes n}$
contains in its decomposition reducible but not fully reducible
representations, which pair up representations that would be distinct
irreducible ones for $q$ generic adding up their $q$-dimensions to
zero. We get pairs of representations that mix in larger structures (type I)
$(V(2{j_k}),V({2j_{k-1}}))$; $(V(2{j_{k-2}}),~V({2j_{k-3}}))$;
$\cdots$.
There
\begin{equation}
\{j_k>j_{k-1}>\cdots>j_1,~~~~0\leq j_1<\frac{1}{2}(p-1)\}~,
\end{equation}
which are related by $j_i=j_1{\rm mod}p$ (if $i$ is odd), and $j_i=p-1-j_1{\rm
mod}p$
(if $i$ is even). Depending on the number of $V's$ we end up with a certain
number of $V({2j_1})$ that cannot mixing and are still irreducible
highest weight representations (type II).
Type-II representations are described by their highest weight vector
$\mid a_j\rangle$ such that $k\mid a_j\rangle=q^j\mid a_j\rangle$ ($0\leq
j<\displaystyle\frac{1}{2}p-1)$. The number of type-II representations reads
\begin{equation}
\Omega_j^{(n)}=\Gamma_j^{(n)}-\Gamma_{p-1-j}^{(n)}+\Gamma_{j+p}^{(n)}
-\Gamma_{p-1-j+p}^{(n)}+\cdots~.
\end{equation}
A special situation occurs if
$j_1=\displaystyle\frac{1}{2}(p-1)$ since $p-1-j_1=j_1$. In this case
the representations $V(2{j_k})$ are still irreducible and do not pair.
Thus the configuration space of the integrable lattice models splits into:
type-I representations which have $q$-dimension zero, and are either
mixed or of the kind $V({(np-1)})$,
and type-II representations which have a nonzero $q$-dimension, and are
still isomorphic to $U\left(su(2)\right)$ ones.
\subsection{Solutions}
Now we are in the position to discuss the solutions of the integrable lattice
models.
Although the XXZ Hamiltonian (\ref{3.0039}) and (\ref{3.0040}) is not
Hermitian, the eigenenergies
are real, because it is invariant under complex conjugation and reflection
symmetry
(relabelling sites from right to left). Because the Hamiltonian has $SU_q(2)$
quantum symmetry, each eigenspace corresponds to an irreducible representation
denoted by a Young pattern. Denote a state with $m$ down spins by
$$\mid x_1,x_2,\cdots,x_m\rangle~,$$
where $x_i$'s are the locations of the down spins on the chain. It is
well-known
that if the total spin of
the chain is $S=\displaystyle\frac{1}{2}(n+1)-m$, the state $\mid
x_1,x_2,\cdots,
x_m\rangle$ are highest weight states. The number of the highest weight
states are given by
\begin{equation}
M_n=\left(\begin{array}{c}
n\\
m\end{array}\right)-\left(\begin{array}{c}
n\\
m-2\end{array}\right)~. \end{equation}
Then, the eigenstates of the Hamiltonian are
\begin{equation}
\mid m,h\rangle=\sum_{(x)}f_h(x_1,x_2,\cdots,x_m)\mid
x_1,x_2,\cdots,x_m\rangle~.
\end{equation}
The highest weight state $\mid m,h\rangle$ of the XXZ spin chain with $m$
down spins can be constructed by the Young operators $Y_m$. For
quantum group $SU_q(2)$ only two-row Young pattern $(n+1-m, m)$ is relevant.
For
each Young pattern we only need one explicit form of the quantum Young
operator corresponding to a Young tableau. To construct the Young operators
explicitly, we first introduce the operators $Z_m^{2m+l}$
\cite{Levy}--\cite{Hou91} as
\begin{equation}
\begin{array}{lcl}
Z^{2m}_m &=& 1~,\\
Z^{2m+1}_m&=&g_{2m+1}(1)~,\\
Z^{2m+2}_m&=&Z^{2m+1}_m g_{2m+2}(2)Z^{2m+1}_m~,\\
\vdots &\vdots& \vdots~,\\
Z^{2m+l}_m&=&Z^{2m+l-1}_m g_{2m+l}(l)Z^{2m+l-1}_m~,\\
\vdots &\vdots& \vdots~,\\
Z^n_m &=& Z^{n-1}_m g_n (n-2m)Z^{n}_m~,
\end{array}
\end{equation}
where we have used the notation $g_i(u)=[1+u]-[u] e_i$. It is easy to see that
the operators $Z^{2m+l}_m$ are constructed by $\{e_i|2m<i\leq 2m+l\}$. It
is also not difficult to prove inductively that the following formulas
for the operators $Z^{2m+l}_m$ are satisfied,
\begin{equation}
e_jZ^{2m+l}_m=Z^{2m+l}_me_j=0~,~~~~\forall 2m<j\leq 2m+l~.
\end{equation}
Let
\begin{equation}
S_m=e_1e_3\cdots e_{2m-1}~,
\end{equation}
then $S_m$ and $Z^{2m+l}_m$ are commutative. Thus we may define the quantum
Young
operators as
\begin{equation}
Y_m=S_mZ_m^n=Z_m^nS_m~.
\end{equation}
It is straightforward to prove that
\begin{equation}
\begin{array}{rcl}
& &Z^n_m Z^n_m\propto Z^n_m~,\\
& &S_m w S_m\propto S_m~,~~~ \forall w \in A_{2m}~,\\
& &Y_m w Y_m\propto Y_m~,~~~ \forall w \in A_n~,\\
& &Y_m Y_m\not=0~.
\end{array}
\end{equation}
Therefore, $Y_m$ is a primitive idempotent, and $wY_m~ (w \in A_n)$
is a primitive left ideal. The different Young patterns describe the
inequivalent
irreducible representations. As a result, the primitive left ideal is
$M_n$ dimensional. Let
\begin{equation}
C^{i_2}_{i_1} = e_{i_2} e_{i_2 -1} \cdots e_{i_1}~,~~~~
\left(i_2 \geq i_1 \right)~,
\end{equation}
then the $M_n$ basis vectors of the primitive ideals are given by
\begin{equation}
\begin{array}{rcl}
& &{^mC_{(i)}}Z^n_m~,\\
& &{^mC_{(i)}}\equiv {^mC_{i_1i_2\cdots i_m}} \equiv C^{i_1}_1 C^{i_2}_3
\cdots C^{i_m}_{2m-1}~,\\
& &1 \leq i_1 <i_2<\cdots<i_m\leq n~,~~ i_k \geq 2k-1~.
\end{array}
\end{equation}
{}From the algebra relations of the Temperley-Lieb algebra, we get the
actions of
$e_i$'s on the basis vectors, ${^mC_{(i)}}Z^n_m$, of the primitive left ideal
as
\begin{itemize}
\item For $t=i_k>i_{k-1}+1$~,
\begin{equation}\label{3.dell1}
e_t\cdot{^mC_{(i)}}Z^n_m=[2]~{^mC_{(i)}}Z^n_m~.
\end{equation}
\item For $t=i_{k+1}=i_k+1,~i_{k-\alpha}\geq t-2\alpha~(1\leq \alpha\leq l),~
i_{k-l-1}<t-2l-2$,
\begin{equation}\label{3.dell2}
e_t\cdot{^mC_{(i)}}Z^n_m={^mC_{(i')}}Z^n_m~,
\end{equation}
where $$\begin{array}{rcl}
& &i_p'=i_p~,~~~~\forall p>k~{\rm or}~p<k-l~,\\
& &i'_{k-\alpha+1}=i_{k-\alpha}~,~~~~1\leq\alpha \leq l~,\\
& &i'_{k-l}=t-2l-2~.
\end{array}$$
\item For $t=i_k+1<i_{k+1}$,
\begin{equation}\label{3.dell3}
e_t\cdot{^mC_{(i)}}Z^n_m={^mC_{(i')}}Z^n_m~,
\end{equation}
where $$\begin{array}{rcl}
& &i_p'=i_p~,~~~~\forall p\not= k~,\\
& &i_k'=i_k+1~.
\end{array}$$
\item For $i_m+1<t\leq n$,
\begin{equation}\label{3.dell4}
e_t\cdot{^mC_{(i)}}Z^n_m=0~.
\end{equation}
\item For $i_{k-1}+1<t<i_k-1,~i_{k+\alpha-1}\geq t+2\alpha~(1\leq\alpha\leq
l),~
i_{k+l}=t+2l+1$,
\begin{equation}\label{3.dell5}
e_t\cdot{^mC_{(i)}}Z^n_m={^mC_{(i')}}Z^n_m~,
\end{equation}
where $$\begin{array}{rcl}
& &i_p'=i_p~,~~~~\forall p<k,~{\rm or}~p>k+l~,\\
& &i_{k+\alpha}'=i_{k+\alpha-1}~,~~~~1\leq\alpha\leq l~,\\
& &i_k'=t~.
\end{array}$$
\item For $t=i_k-1>i_{k-1}+1$,
\begin{equation}\label{3.dell6}
e_t\cdot{^mC_{(i)}}Z^n_m={^mC_{(i')}}Z^n_m~,
\end{equation}
where $$\begin{array}{rcl}
& &i_p'=i_p~,~~~~\forall p\not=k~,\\
& &i_k'=i_k-1~.
\end{array}$$
\item For $i_{k-1}+1<t<i_k-1,~i_{k+\alpha-1}\geq t+2\alpha~~(1\leq\alpha
\leq m-k+1)$,
\begin{equation}\label{3.dell7}
e_t\cdot{^mC_{(i)}}Z^n_m=0~.
\end{equation} \end{itemize}
It is obvious from
Eq.(\ref{3.I212}) that the number of down spins of a state which are not
annihilated by the Young operator ${^mC_{(i)}}Z^n_m$ must be not less than
$m$, and a state of the highest weight of the representation has $m$ down
spins located at the first $2m$ positions. We construct the eigenspace of
the XXZ Hamiltonian by acting the basis of the primitive left ideal,
${^mC_{(i)}} Z_m^n$, on the state $\mid x_1,x_2,\cdots,x_m\rangle$. From
the definition relation of the elements of Temperley-Lieb algebra, we know
that if both $i$-th and $i+1$-th spins are down, the action of $e_i$
returns zero. The Young operators $Y_m$ contain $S_m~(e_1e_3\cdots
e_{2m-1})$, and the direction of any spin among the first $2m$ spins is
different from its neighbor's; the left $n-2m$ spins are all up. Thus the
action of $Z^n_m$ is a constant. For definiteness we define the state with
the highest weight as follows
\begin{equation}
{^m\Psi}_{(i)s}\equiv {^mC}_{(i)}\mid 1,3,5,\cdots,(2m-1)\rangle~,~~~
s=\frac{1}{2}(n+1-2m)~,
\end{equation}
where the constant factor obtained by applying $Z^n_m$ on the state
$\mid 1,2,3,\cdots,(2m-1)\rangle$ has been neglected, the subscript $s$ is
the highest weight eigenvalue of $\Delta^{(n)} \left(s_z \right)$.
Since the different Young patterns describe the inequivalent irreducible
representations, the corresponding energies are generally
different. However, for the same Young pattern $(n-m,~ m)$
there are $M_n$ Young operators ${^mC}_{(i)}Z^n_m$ and $M_n$ spaces with
the same irreducible representation.
The linear combinations of ${^m\Psi}_{(i)s}$ with the same $m$ form the
eigenstates of the XXZ Hamiltonian
\begin{equation}\label{3.dell8}
\begin{array}{rcl}
& &H\cdot{^m\Phi}_{hs}={^mE}_h ~{^m\Phi}_{hs}~, \\
& &{^m\Phi}_{hs}=\displaystyle\sum_{(i)}a^h_{(i)}~ {^m\Psi}_{(i)s}~,
~~~h=1,~2,~\cdots,~M_n.
\end{array}
\end{equation}
Using the lowering operator $\Delta^{(n)}(s^-)$, we get easily the partners
${^m\Phi}_{h\alpha}$ in the representation
\begin{equation}
{^m\Phi}_{h\alpha}=[s-\alpha]^{-1}\Delta^{(n)}(s^-)~ {^m\Phi}_{h(\alpha+1)}~.
\end{equation}
Making use of Eqs.(\ref{3.dell1})---(\ref{3.dell7}), we transform
Eq.(\ref{3.dell8}) to be
a set of coupled linear algebraic equations with respect to
$a^h_{(i)}$, which can be solved by the standard method.
We now discuss a simple example in detail. For $m=0$, it is obvious that there
is only one state $\mid~\rangle_0$ with all spins up and zero eigenenergy.
For $m=1$, the simplest nontrivial case, the equations of motion of
the XXZ system are
\begin{equation}
\begin{array}{rcl}
& &{^1\Phi}_{h\frac{n-1}{2}}=\displaystyle\sum_ia_i^h~{^1C}_i\mid1\rangle~,\\
& &\left([2]-{^1E}_h\right)a_i^h+a_{i+1}^h+a_{i-1}^h=0~,\\
& &a_0^h=a_{n+1}^h=0~.
\end{array}
\end{equation}
By induction, we can easily prove that
\begin{equation}
\frac{a_i^h}{a_1^h}=(-1)^{i-1}\frac{\sin(ih\pi/(n+1))}{\sin(h\pi/(n+1))}~.
\end{equation}
For simplicity, we put $a_1^h=1$, and the eigenenergy is calculated from
$a_{n+1}^h=0$~,
\begin{equation}
{^1E}_h=[2]-2\cos\left(\frac{h\pi}{n+1}\right)~,~~~~h=1,~2,~\cdots,~n.
\end{equation}
\section{Conformal field theory}
During the past ten years, two-dimensional conformal invariant quantum
field theory has become of importance in statistical physics
and string theory \cite{Belavin84}--\cite{Kaku},\cite{Cardy87}.
Conformal field theory
(CFT) turned out due to be relevant in statistical models
exhibiting a second order phase transition. At such a phase
transition the typical configurations have fluctuations on all length scales,
so that the field theory describing the model at a critical point is
expected to be invariant under scale transformations. Unlike in three
and higher dimensions the conformal group is infinite-dimensional in two
dimensions. This fact imposes significant constraints on two-dimensional
CFT's. Ultimately, we may hope that these
constraints are strong enough to obtain a complete classification of
two-dimensional critical systems. CFT's are also essential
to string theory, which assume that the
elementary particles are not point-like but rather
behave as one dimensional objects, called strings. In perturbation theory,
the amplitude of a string can be expressed as sums over all possible
two-dimensional world-sheets of various topology, swept out by a string in
space-time. The terms in perturbation series only depend on the conformal
equivalence class of the two-dimensional metric on the Riemann surface.
A CFT is characterized by its scale invariance \cite{Belavin84,Friedan84}.
As a local field
theory, scale invariance implies full conformal invariance. Scale invariance
is equivalent to the vanishing trace of the energy-momentum tensor.
In complex coordinates this means that $T_{z\bar{z}} = 0$. Thus,
there are only two independent components of the energy-momentum tensor
$T_{zz},~T_{\bar{z}\bar{z}}$ and their conservation law implies
that
$T_{zz} \equiv T(z)$ $(T_{\bar{z} \bar{z}} \equiv
T(\bar{z}))$ is a holomophic (antiholomorphic) function of
$z$ $(\bar{z})$. The generators of infinitesimal conformal transformations
are
\begin{equation}\label{4.2.1}
L_n = \oint_C \frac{dz}{2 \pi i} z^{n+1} T(z)~,~~~~(n\in Z)~,
\end{equation}
and similarly for $\overline{L}_n$. The contour circles the origin only once.
Its
shape do not matter as a consequence of the conservation law and Cauchy's
theorem. By making use of Eq.(\ref{4.2.1}), we can write
\begin{equation}
T(z) = \sum_{n \in Z} L_n z^{-n-2}~,~~~~
\overline{T} (\bar{z}) = \sum_{n \in Z} \overline{L}_n\bar{z}^{-n-2}~.
\end{equation}
Out of all fields in CFT, primary fields behave as
$(h,\overline{h})$ tensors, i.e.
\begin{equation}
\Phi_{h,\bar{h}} (z,\bar{z}) d z^h d \bar{z}^{\bar{h}}
\end{equation}
is invariant under conformal transformations. $(h,\bar{h})$ are called the
conformal dimensions of the primary fields. By analytical properties of
$T(z)$ the correlation function
\begin{equation}
\langle T (z) \Phi_1 (z_1) \cdots \Phi_N (z_N) \rangle~,
\end{equation}
can be evaluated using the operator product expansion (OPE)
\begin{equation}\label{4.2.5}
T (z) \Phi (w) = \frac{h}{(z-w)^2} \Phi(w) + \frac{1}{z-w}
\partial_w \Phi(w) + \cdots~.
\end{equation}
As a function of $z,~ T(z)$ is a meromorphic quadratic differential
$(h = 2)$ on the sphere. The only singularities appear at points
$z_1,~ z_2,~ \cdots,~ z_N$ and they are determined by the OPE (\ref{4.2.5}),
\begin{equation}\label{4.2.6}
\langle T (z) \Phi_1 (z_1) \cdots \Phi_N (z_N)\rangle
= \sum^N_{j=1} \left( \frac{h_i}{(z-z_j)^2}
+ \frac{1}{z-z_j} \frac{\partial}{\partial z_j} \right)
\langle\Phi_1(z_1) \cdots \Phi_N (z_N) \rangle~.
\end{equation}
The OPE between two $T$'s is given by
\begin{equation}
T(z) T(w) = \frac{{c}/{2}}{(z-w)^4}+
\frac{2}{(z-w)^2} T(w)+\frac{1}{z-w}
\partial_w T(w) +\cdots~.
\end{equation}
This implies that the generators of infinitesimal conformal
transformations, $L_{n}$'s satisfy the following commutators
\begin{equation}
[L_n,L_m] = (n - m) L_{n+m}+ \frac{c}{12} (n^3 - n) \delta_{n+m,0}~,
\end{equation}
and we obtain the Virasoro algebra. The symmetry of a CFT under the
Virasoro algebra Vir$\bigotimes\overline{\rm Vir}$ means that the fields of the
theory fall into different conformal families, each generated by a
primary field. The states generated by the action of $T(z)$ on the
primary field, known as descendent fields, are determined by those of
the primary fields through the use of the conformal Ward identity
(\ref{4.2.6}).
\subsection{Minimal model}
The minimal model in a sense, is the simplest CFT. The Hilbert space of
the theory is a finite sum of irreducible highest weight representation
space of the Virasoro algebra. There are finitely many primary fields, and
correlation functions of conformal fields are expressed in terms of a
finite sum of
holomorphically factorized terms. In principle, conformal invariance allows
one to compute all matrix elements of conformal fields in terms of
finitely many structure constants of the theory. And the choice
of structure
constants are restrained by the requirement of locality. Imposing this
requirement, one can in principle compute the possible values of the
structure constants.
The Coulomb gas version of the minimal model is defined in terms of
a massless scalar field $\phi$ in the presence of a background charge
$2 \alpha_0$ located at infinity \cite{Dotsenko84,Dotsenko85}. The
corresponding energy-momentum tensor
defines representations of the Virasoro algebra on the Fock modules ${\cal
F}_\alpha$,
\begin{equation}
\begin{array}{rcl}
& & T(z) = -\frac{1}{4} \partial \phi(z) \partial
\phi(z) + i \alpha_0 \partial^2 \phi(z)~,\\
& &\langle\phi(z) \phi(w)\rangle= -2\ln (z - w)~,\\
& &c = 1 - 24\alpha^2_0~.
\end{array}
\end{equation}
The highest weight vector of ${\cal F}_\alpha$ is given by the vertex operator
\begin{equation}
V_\alpha(z)=:e^{i\alpha\phi(z)}:~
\end{equation}
with $U(1)$ charge $\alpha$.
$T(z)$ follows from variation of the action for a free scalar field
with some background charge: a term proportional to $R \phi$ in the
Lagrangian where $R$ is the two-dimensional scalar curvature. Thus the
$U(1)$ current $\partial \phi(z)$ is anomalous
\begin{equation}
T(z) \partial \phi(w) = \frac{1}{(z-w)^2} \partial \phi(w) +
\frac{1}{z-w} \partial^2 \phi(w) +
\frac{4i \alpha_0}{(z-w)^3}+\cdots~.
\end{equation}
The last term represents the anomalous. The background charge changes the
conformal dimension for the case $\alpha_0 = 0$
\begin{equation}
T(z):e^{i\alpha\phi(w)}:=\frac{\alpha(\alpha-2\alpha_0)}{(z-w)^2}:
e^{i\alpha\phi(w)}:+\frac{1}{z-w}\partial_w:
e^{i\alpha\phi(w)}:+\cdots~.
\end{equation}
The conformal dimension of the vertex operator,
$\Delta_d\left(V_{\alpha}(z)\right)=\alpha(\alpha-2\alpha_0)$ and the
correlation
functions of these vertex operators will vanish unless the background
charge is screened
\begin{equation}
\langle V_{\alpha_1}(z_1)V_{\alpha_2}(z_2)\cdots V_{\alpha_N}(z_N)\rangle
=\left\{\begin{array}{l}
\displaystyle\prod_{i<j}(z_i-z_j)^{2\alpha_i
\alpha_j}~,~~~{\rm if}~
\sum_{i=1}^N\alpha_i=2\alpha_0~,\\
0~,~~~ {\rm otherwise}~.\end{array}
\right.
\end{equation}
The braid relation satisfied by the Coulomb gas vertex operators is
\begin{equation}\label{4.2.14}
V_{\alpha_1}(z_1)V_{\alpha_2}(z_2)
= e^{2\pi i\alpha_1\alpha_2} V_{\alpha_2}
(z_2) V_{\alpha_1}(z_1)~.
\end{equation}
For $\alpha=\alpha_\pm$, where $2\alpha_0=\alpha_++\alpha_-$ and
$\alpha_+\alpha_- = -1$, we obtain the screening operators $J(z)=V_
{\alpha_\pm}(z)$ with conformal dimension equal to one. It is
important to note that both $V_{\alpha}(z)$ and its dual
$\tilde{V}_\alpha(z)=V_{2\alpha_0-\alpha}
(z)$ have the same conformal dimension. In particular, we can write
the identity in two ways, $1$ and $:\exp\left(2i\alpha_0\phi\right):$.
The existence
of currents of conformal dimension one allows us to introduce screening
operator $Q$,
\begin{equation}
Q = \int_\Gamma J(z)dz~,
\end{equation}
where the contour $\Gamma$ is chosen so that $Q$ acts as an intertwiner,
i.e., without
affecting the conformal properties of the correlation function. For
example if $z_1,~z_2$ are the insertion points of vertex operators, the
screening charge
\begin{equation}
\int^{z_2}_{z_1} dzJ_i(z)
\end{equation}
is an intertwiner. Inserting $Q$ in correlation functions does not change
the conformal properties. If we write down the $4$-point function of an
operator $V_\alpha(z)$, $\langle V_\alpha V_\alpha V_\alpha
\tilde{V}_\alpha\rangle$.
The total charge is $3\alpha +2\alpha_0-\alpha=2\alpha+2\alpha_0$, hence
will vanish generally unless
\begin{equation}
2\alpha = -n\alpha_+-m\alpha_-~.
\end{equation}
In the case we can introduce $nQ_+$ and $mQ_-$ screening operators
to obtain a nonvanishing amplitude. Therefore, the spectrum of vertex
operators with nonvanishing 4-point functions is given by
\begin{equation}\label{4.2.18}
\alpha_{m,n}= \frac{1-n}{2}\alpha_+ +\frac{1-m}{2}\alpha_-~,~~~~
n,~m\geq1~.
\end{equation}
A $4$-point block hence takes the form
\begin{equation}
\begin{array}{rcl}
& &\displaystyle\int_{\Gamma_1}dt_1\int_{\Gamma_2}dt_2\cdots
\int_{\Gamma_n}dt_n \int_{\Gamma'_1} dt'_1 \int_{\Gamma'_2} dt'_2
\cdots \int_{\Gamma'_m} dt'_m\times\\
& &\times\langle V_{\alpha_1}(z_1)V_{\alpha_2}(z_2)V_{\alpha_3}(z_3)
V_{\alpha_4}(z_4)J_+(t_1) J_+(t_2)\cdots J_+(t_n)J_-(t'_1) J_-(t'_2)
\cdots J_-(t'_m) \rangle~.
\end{array}
\end{equation}
The general correlation function in the Coulomb gas representation is\\ \\
$\langle V_{\alpha_1}(z_1) V_{\alpha_2}(z_2)\cdots V_{\alpha_N}(z_N)\rangle$
\begin{equation}
\sim\langle V_{\alpha_1}(z_1)Q_i(z_a,~z_b)\cdots Q_j(z_c,~z_d)V_{\alpha_2}
(z_2) \cdots Q_k(z_e,~z_f) V_{\alpha_N}(z_N) \rangle~,
\end{equation}
where $z_a,~z_b,~\cdots,~z_e,~z_f \in \{z_1,~z_2,~\cdots,~z_N\}$,
and total charge equal to zero.
We define the screening operators $X^-_\pm$ \cite{Gomez90,Gomez91} by
\begin{equation}
X^-_{\pm}V_\alpha(z)=\frac{1}{1-q^{-1}_\pm} \int_G dtJ_\pm V_\alpha(z) \sim
Q(\infty,~z) V_\alpha(z)~,
\end{equation}
where $q_\pm=e^{2\pi i\alpha^2_\pm}$, and the contour $G$ surrounds the cut
from $z$ to infinity (Fig.19). Owing to the $SL(2,C)$ invariance of the
correlation function, we can write the $N$-point conformal block as
\begin{equation}
\lim_{z_1\to\infty}\langle \tilde{V}_{\alpha_1}(z_1)X^-_+ \cdots
X^-_-V_{\alpha_2} \cdots X^-_+
X^-_-\cdots V_{\alpha_N}(z_N) \rangle~,
\end{equation}
with vanishing total Coulomb charge. The screening operator
$X^-_\pm$'s are understood to act only on the first vertex operator
to their right.\\
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,23)(0,4)
\put(35,10){\oval(8,5)[b]}
\put(31,10){\line(0,1){15}}
\put(39,10){\line(0,1){15}}
\put(31,25){\vector(0,-1){7.5}}
\put(39,10){\vector(0,1){7.5}}
\multiput(35,10.5)(0,1.5){10}{\line(0,1){1}}
\put(34.5,10){$\bullet$}
\put(36,9.4){$z$}
\put(40,17){$G$}
\end{picture}
\hspace{7.5em}Fig.19. The defining contour of generators
\hspace{10.7em}of quantum group.
Introduce the operator $k_\pm$ by
\begin{equation}
k_\pm V_\alpha(z) =\exp\left(-2i\pi\alpha_\pm \oint_c \partial\phi \right)
V_\alpha(z)~.
\end{equation}
The Borel subalgebra, generated by the operators $X^-_\pm$ and
$k_\pm$, of the quantum group underlying the minimal model is
\begin{equation}
\begin{array}{rcl}
[X^-_+,X^-_-] &=& 0,~~~~~[k_+,~k_-] =0~,\\
k_\pm X^-_\pm k^{-1}_\pm &=& q^{-1}_\pm X^-_\pm,~~~
k_\pm X^-_\mp k_\pm^{-1}=-X^-_\mp~.
\end{array}
\end{equation}
To define the dual operators of $X^-_\pm$, $X^+_\pm$, we shall make use of the
transformation law of the screened vertex operator under the Virasoro
algebra. Explicitly, we define $X^+_\pm$ by\\ \\
$\delta_\xi \left((X^-_+)^n (X^-_-)^mV_\alpha(z)\right) - (X^-_+)^n(X^-_-)^m
\left(\delta_\xi V_\alpha(z)\right)$
\begin{equation}\label{4.2.29}
\begin{array}{rcl}
&=&-(1-q_+^{-1})\xi(\infty)J_+(\infty)X^+_+(X^-_+)^n(X^-_-)^mV_\alpha(z)\\[2mm]
& &-(1-q_-^{-1})\xi(\infty)J_-(\infty)X^+_-(X^-_+)^n(X^-_-)^mV_\alpha(z)~,
\end{array}
\end{equation}
where $\xi(z)$ is the vector field that generates the conformal
transformation. Making use of the commutator between the Virasoro
generator $L_n$ and the screening operator $J_\pm$,
\begin{equation}
[L_n,J_\pm] = \frac{d}{dz}\left( z^{n+1}J_\pm(z)\right)~,
\end{equation}
we get\\ \\
$L_n(X^-_+)^n(X^-_-)^mV_\alpha(z) $
\begin{equation}\label{4.2.31}
\begin{array}{rcl}
&=&(X^-_+)^n(X^-_-)^m\left(L_nV_\alpha(z)\right)\\[3mm]
& &-\displaystyle\lim_{t\to\infty}
t^{n+1}J_+(t)|[n]_{q_+^{-1}}(1-e^{4\pi
i\alpha\alpha_+}q^{n-1}_+)(X^-_+)
^{n-1}(X^-_-)^mV_\alpha(z)\\[3mm]
& &-\displaystyle\lim_{t\to\infty}
t^{n+1}J_-(t)|[m]_{q_-^{-1}}(1-e^{4\pi i
\alpha\alpha_-}q^{m-1}_-)(X^-_+)^n(X^-_-)^{m-1}V_\alpha(z)~,
\end{array}
\end{equation}
where $\displaystyle|[x]_q=\frac{1-q^{-x}}{1-q^{-1}}$.\\
{}From Eqs.(\ref{4.2.29}) and (\ref{4.2.31}) we obtain that
\begin{equation}\label{4.2.32}
\begin{array}{rcl}
X^+_+(X^-_+)^n(X^-_-)^mV_\alpha(z) &=&
|[n]_{q_+^{-1}}\displaystyle\frac{1-e^{4\pi i \alpha\alpha_+}
q^{n-1}_+}{1-q^{-1}_+}(X^-_+)^{n-1}(X^-_-)^mV_\alpha(z)~,\\[3mm]
X^+_-(X^-_+)^n(X^-_-)^mV_\alpha(z) &=&
|[m]_{q_-^{-1}}\displaystyle\frac{1-e^{4\pi i \alpha\alpha_-}
q^{m-1}_-}{1-q^{-1}_-}(X^-_+)^n(X^-_-)^{m-1}V_\alpha(z)~.
\end{array}
\end{equation}
{}From Eqs.(\ref{4.2.29}) and (\ref{4.2.32}) one can verify that
$\delta_\xi$ satisfies
the Virasoro algebra, $X^+_+$ commutes with $X^+_-$, and $X^+_\pm$ commute
with the Virasoro algebra. Now we can write down the other Borel
subalgebra of the quantum group underlying the minimal model as
\begin{equation}
\begin{array}{rcl}
& &[X^+_+,X^+_-]=0~,\\
& &k_\pm X^+_\pm k_\pm^{-1}=q_\pm X^+_\pm~,~~~~
k_\pm X^+_\mp k_\pm^{-1}=-X^+_\mp~,\\
& &X^+_\pm X^-_\pm - q_\pm X^-_\pm X^+_\pm=\displaystyle\frac{1-k^{-2}_\pm}
{1-q^{-1}_\pm}~,\\
& &X^+_\pm X^-_\mp-q_\pm X^-_\mp X^+_\pm=0~.
\end{array}
\end{equation}
Let us define the action of the operators $X^-_\pm$ on the space for
the ordinary operator product $V_\alpha(z_1)V_\beta
(z_2)$ corresponding to the comultiplication operation
of Hopf algebra
\begin{equation}
\Delta\left( X^-_\pm\right) \left( V_\alpha(z_1)V_\beta(z_2) \right) =
\int_{\Delta G}dtJ_\pm(t)V_\alpha(z_1)V_\beta(z_2)~,
\end{equation}
where the contour $\Delta G$ surrounds the operator $V_\alpha(z_1)
V_\beta(z_2)$ (Fig.20). Deforming the contour $\Delta G$ into the union
$G_1\cup G_2$ with $G_i$ ($i=1,~2$) the contour surrounding the vertex
at point $z_i$, we obtain
\begin{equation}\label{4.2.35}
\Delta \left(X^-_\pm\right)\left(V_\alpha(z_1)V_\beta(z_2)\right) =
X^-_\pm\left(V_\alpha(z_1)\right)
V_\beta(z_2) +
k^{-1}_\pm\left(V_\alpha(z_1)\right)X^-_\pm\left(V_\beta(z_2)\right)
{}~.
\end{equation}
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,23)(0,4)
\put(10,10){\oval(15,5)[b]}
\put(2.5,10){\line(0,1){15}}
\put(17.5,10){\line(0,1){15}}
\put(2.5,25){\vector(0,-1){9.5}}
\put(17.5,10){\vector(0,1){7.5}}
\multiput(7.5,10.5)(0,1.5){10}{\line(0,1){1}}
\multiput(12.5,10.5)(0,1.5){10}{\line(0,1){1}}
\put(7,10){$\bullet$}
\put(8.5,9.4){$z_1$}
\put(12,10){$\bullet$}
\put(13.5,9.4){$z_2$}
\put(8,5){$\Delta G$}
\put(21.5,15){$=$}
\put(35,10){\oval(8,5)[b]}
\put(31,10){\line(0,1){15}}
\put(39,10){\line(0,1){15}}
\put(31,25){\vector(0,-1){9.5}}
\put(39,10){\vector(0,1){7.5}}
\multiput(35,10.5)(0,1.5){10}{\line(0,1){1}}
\multiput(44,10.5)(0,1.5){10}{\line(0,1){1}}
\put(34.5,10){$\bullet$}
\put(36,9.4){$z_1$}
\put(43.5,10){$\bullet$}
\put(45,9.4){$z_2$}
\put(34,5){$G_1$}
\put(50,15){$+e^{2\pi i\alpha_\pm\alpha}$}
\put(69,10){\oval(8,5)[b]}
\put(65,10){\line(0,1){15}}
\put(73,10){\line(0,1){15}}
\put(65,25){\vector(0,-1){9.5}}
\put(73,10){\vector(0,1){7.5}}
\multiput(69,10.5)(0,1.5){10}{\line(0,1){1}}
\multiput(60,10.5)(0,1.5){10}{\line(0,1){1}}
\put(59.5,10){$\bullet$}
\put(61,9.4){$z_1$}
\put(68.5,10){$\bullet$}
\put(70,9.4){$z_2$}
\put(69,5){$G_2$}
\end{picture}
\hspace{7.5em}Fig.20. Comultiplication operation of $X_\pm^-$.
The diagonal operators $k_\pm$ in Eq.(\ref{4.2.35}) arises
from the braiding between the screening operator $J_\pm$ and the vertex
operator $V_\alpha(z)$. The comultiplication operation for $X^-_\pm$
follows easily from Eq.(\ref{4.2.35}), taking form
\begin{equation}
\Delta( X^-_\pm) =X^-_\pm \otimes 1 + k^{-1}_\pm \otimes X^-_\pm~.
\end{equation}
Similarly the comultiplication operation for $k_\pm$ is given by
\begin{equation}
\Delta (k_\pm) = k_\pm \otimes k_\pm~.
\end{equation}
The comultiplication operation for $X^+_\pm$ follows from that for the
Virasoro operators, \\ \\
$\delta_\xi
\left((X^-_+)^n(X^-_-)^mV_\alpha(z_1)(X^-_+)^{n'}(X^-_-)^{m'}V_\beta(z_2)\right)$
\begin{equation}
\begin{array}{rcl}
&=& \left(\delta_\xi\left( (X^-_+)^n(X^-_-)^mV_\alpha(z_1)\right)\right)
(X^-_+)^{n'}(X^-_-)^{m'}V_\beta(z_2)\\[3mm]
& & +(X^-_+)^{n}(X^-_-)^{m}V_\alpha(z_1) \left(\delta_\xi\left(
(X^-_+)^{n'}(X^-_-)^{m'}V_\beta(z_2)\right)\right)\\[3mm]
&=& (X^-_+)^{n}(X^-_-)^{m}\left(\delta_\xi V_\alpha(z_1)\right)
(X^-_+)^{n'}(X^-_-)^{m'}V_\beta
(z_2)\\[3mm]
& & + (X^-_+)^{n}(X^-_-)^{m}V_\alpha(z_1)
(X^-_+)^{n'}(X^-_-)^{m'}\left(\delta_\xi V_\beta(z_2)\right)\\[3mm]
& & -\left( (1-q^{-1}_+)\xi(\infty)J_+(\infty)\Delta (X^+_+) +
(1-q^{-1}_-)\xi(\infty)J_-(\infty)\Delta(X^+_-) \right)\times\\[3mm]
& & ~~\times (X^-_+)^{n}(X^-_-)^{m}V_\alpha(z_1)(X^-_+)^{n}(X^-_-)^{m}
V_\beta(z_2)~,
\end{array}
\end{equation}
where we have used Eq.(\ref{4.2.29}) and
\begin{equation}
\Delta (X^+_\pm) = X^+_\pm \otimes 1 + k^{-1}_\pm \otimes X^+_\pm~.
\end{equation}
Besides the comultiplication $\Delta$, the other Hopf operations such as
antipode $S$ and counit $\epsilon$ are introduced by the following
formulas
\begin{equation}
\begin{array}{rcl}
\epsilon (k_\pm) &=& 1~,~~~~ \epsilon( X^-_\pm) = 0 = \epsilon X^-_\pm~,\\
S(k_\pm) &=& k^{-1}_\pm~,~~~~S(X^-_\pm)=-k_\pm X^-_\pm~,~~~~S(X^+_\pm)
=-k_\pm X^+_\pm~.
\end{array}
\end{equation}
The antipode $S$ is essentially a path reversing operator, while
the counit is a contour killing mapping (i.e. $\epsilon\left(\displaystyle\oint
_CX\right)=0$ for any $X$). The quantum group generated by
$k_\pm,~X^-_\pm$ and $X^+_\pm$ summarizes the quantum symmetries of
the minimal model. For the thermal subalgebra
$\alpha_{n,1}$ (or $\alpha_{1,m})$, the operators $k_+,~X^-_+$
and $X^+_+$ (or $k_-,~X^-_-$ and $X^+_-)$ generate the quantum group
$SU_q(2)$.
For each vertex operator $V_\alpha(z)$ is associated with a representation
space
${\cal V}^\alpha$ generated by the screened vertex operator $e^\alpha_{
n_+,n_-}$,
\begin{equation}\label{4.2.41}
e^\alpha_{n_+,n_-} = (X^-_+)^{n_+}(X^-_-)^{n_-}V_\alpha(z)~.
\end{equation}
The screened vertex operators $e_{n_+,n_-}^\alpha$ give the representations of
the quantum group on the space ${\cal V}^\alpha$
\begin{equation}\label{4.2.42}
\begin{array}{rcl}
& & k_\pm e^\alpha_{n+,n-}=\exp\left(-2\pi i\alpha_\pm(\alpha+n_+\alpha_++n_-
\alpha_-)\right)e^\alpha_{n_+,n_-}~,\\[3mm]
& & X^-_+e^\alpha_{n_+,n_-}=e^\alpha_{n_++1,n_-}~,~~~~
X^-_-e^\alpha_{n_+,n_-} = e^\alpha_{n_+,n_- +1}~,\\[3mm]
& & X^+_+e^\alpha_{n_+,n_-}=|[n_+]_{q_+^{-1}}\displaystyle\frac{1-e^{4
\pi i\alpha\alpha_+}q^{n_+-1}_+}{1-q^{-1}_+}
e^\alpha_{n_+ -1,n_-}~,\\[3mm]
& & X^+_-e^\alpha_{n_+,n_-}=|[n_-]_{q_-^{-1}}\displaystyle\frac{1-e^{4\pi
i \alpha\alpha_-}q^{n_- ^{-1}}}{1-q^{-1}_-}
e^\alpha_{n_+,n_- -1}~.
\end{array}
\end{equation}
The space ${\cal V}^\alpha$ will be finite-dimensional
provided there
exist positive integers $n^+_\alpha$ and $n^-_\alpha$ such that
$e^\alpha_{n^+_\alpha,n^-_\alpha}=0$. In this case it is easy
to see from Eq.(\ref{4.2.41}) that the charge $\alpha$ is given
by the Kac's formula (\ref{4.2.18}), with $n=n^+_\alpha$ and $m=n^-_\alpha$.
The dimension of
${\cal V}^\alpha$ is therefore equals to $n^+_\alpha n^-_\alpha = nm$.
In a rational theory, where $\alpha^2_+=p'/p$, the range of the
positive integers $n$ and $m$ that define a finite-dimensional space
${\cal V}^{\alpha_{n,m}}$, is restricted to the intervals $1\leq n\leq p$
and $1\leq m\leq p'$. This is due to the path ordering of the
screening operator $J_+$ and $J_-$ in Eq.(\ref{4.2.41}).
In order to see
this phenomenon in more detail we will write the screened vertex
operators $e^\alpha_{n_+,n_-}$ alternatively
\begin{equation} \label{4.2.40}
e^\alpha_{n_+,n_-}(z) = \prod^{n_+-1}_{r=0}\left(1-e^{4\pi i
\alpha\alpha_+}q^r_+ \right)
\prod^{n_--1}_{r'=0} \left(1-e^{4\pi i
\alpha\alpha_-}q^{r'}_- \right)
\tilde{e}^\alpha_{n_+,n_-}(z)~,
\end{equation}
where
\begin{equation}
\begin{array}{rcl}
\tilde{e}^\alpha_{n_+,n_-}(z) &=& (P_+)^{n_+}(P_-)^{n_-}V_\alpha(z)\\
&=&\displaystyle\int^z_\infty dt_1J_+(t_1) \cdots
\int^z_\infty dt_{n_+}J_+(t_{n_+})
\int^z_\infty dt'_1J_-(t'_1)\cdots
\int^z_\infty dt'_{n_-}J_-(t'_{n_-})
V_\alpha(z),
\end{array}
\end{equation}
here we have used the notation
\begin{equation}
P_\pm = \int^z_\infty J_\pm(z)dt~.
\end{equation}
It is convenient to path order the integrals entering the definition
of $\tilde{e}^\alpha_{n_+,n_-}(z)$. Denoting the``path-ordering" operator
by the symbol $T$ (i.e. $|t_1|>|t_2|> \cdots>|z|)$ we obtain
\begin{equation}\label{4.2.43}
\tilde{e}^\alpha_{n_+,n_-}(z) = |[n_+]_{q_+^{-1}}!|[n_-]_{q_-^{-1}}!
T(\tilde{e}^\alpha_{m_+,n_-}(z))~.
\end{equation}
Eq.(\ref{4.2.43}) is the $q$-analogue of the path ordering of an integral
operator.
{}From Eqs.(\ref{4.2.40}) and (\ref{4.2.43}), we see that
$e^\alpha_{n_+,n_-}(z)$ is
non-zero, if $1\leq n_+ \leq n^\pm_\alpha-1$, with $1\leq
n=n^+_\alpha
\leq p$ and $1\leq m=n^-_\alpha\leq p'$.
\subsection{WZNW model}
In the previous subsection we have considered the CFT with symmetry algebra
Vir$\otimes\overline {\rm Vir}$. In general, CFT has
a symmetry algebra ${\cal U}$, the chiral algebra ${\cal U}={\cal U}_L\otimes
{\cal U}_R$ with left- and right-handed components.
The chiral algebras are characteristic of ${\cal U}_L$ (${\cal
U}_R$) always
containing the identity operator and the Virasoro ($\overline{\rm Vir}$)
algebra.
Making use of a set of complex fermions $b^i(z)$ and $c_i(z)$
($i=1,~2,~\cdots,~d$), for an algebra $g$ and a real representation
$(T^a)_{ij}$, ($i,~j=1,~2,~\cdots,~d$), we can construct currents
\begin{equation}
J^a(z)=b^i(T^a)_i^jc_j=bT^ac=\sum_{n\in Z}J_n^az^{-n-1}~.
\end{equation}
The nontrivial OPEs are
\begin{equation}
\begin{array}{rcl}
& &b^i(z)c_j(w)=\displaystyle\frac{\delta^i_{j}}{z-w}+\cdots~,\\[2mm]
& &c_i(z)b^j(w)=\displaystyle\frac{\delta_i^{j}}{z-w}+\cdots~,
\end{array}
\end{equation}
and the commutation relations of $J_n^a$ follow from the OPE
\begin{equation}
J^a(z)J^b(w)=\frac{{\rm Tr}T^aT^b}{(z-w)^2}+if^{abc}\frac{J^c(w)}{z-w}+\cdots~.
\end{equation}
Choosing an appropriate basis for the Lie algebra $g$,
\begin{equation}
{\rm Tr}T^aT^b=k\delta^{ab}~,
\end{equation}
we can always write down the commutation relations of the currents algebra as
\begin{equation}
[J_n^a,J_m^b]=if^{abc}J_{n+m}^c+n\delta^{ab}k\delta_{n+m,0}~.
\end{equation}
The operators $J_n^a$'s are associated with the loop algebra of $g$,
\begin{equation}
\epsilon(z)=\sum_{a,n}\epsilon_n^aT^az^n~,
\end{equation}
where $\epsilon_n^a$'s are the generators of the loop algebra, their
commutation relations always generate a Kac-Moody algebra. If the ground state
of a CFT is invariant under the Kac-Moody algebra, the fields of the
theory are catalogued into families, each of which provides an irreducible
representation of the Kac-Moody algebra.
A primary field has the properties
\begin{equation}\label{4.2.50}
J^a(z)\Phi_j(w)=-(T^a)_{jk}\frac{\Phi_k(w)}{z-w}+\cdots~.
\end{equation}
And the primary fields also have the OPEs
\begin{equation}\label{4.2.51}
\begin{array}{rcl}
& & T(z)\Phi_i(w)=\displaystyle\frac{h}{(z-w)^2}\Phi_i(w)+\frac{1}{z-w}\partial
_w\Phi_i(w)+\cdots~,\\[3mm]
& & T(z)J^a(w)=\displaystyle\frac{1}{(z-w)^2}J^a(w)+\frac{1}{z-w}\partial
_wJ^a(w)+\cdots~.
\end{array}
\end{equation}
And then
\begin{equation}
J^a(z)J^b(w)=\frac{k\delta^{ab}}{(z-w)^2}+if^{abc}\frac{J^c(w)}{z-w}
+\cdots~.
\end{equation}
Eqs.(\ref{4.2.50}) and (\ref{4.2.51}) allow us to determine the
correlation functions explicitly, provided that all the fields involved are
primary ones,\\
$\langle T(z)\Phi_{i_1}(z_1)\Phi_{i_2}(z_2) \cdots\Phi_{i_N}(z_N) \rangle$
\begin{equation}
\begin{array}{c}
=\displaystyle
\sum_{j=1}^N\left(\frac{h_j}{(z-z_j)^2}+\frac{1}{z-z_j}\frac{\partial}{\partial
z_j}
\right)\langle\Phi_{i_1}(z_1)\Phi_{i_2}(z_2) \cdots\Phi_{i_N}(z_N)
\rangle~,\\[2mm]
\langle J^a(z)\Phi_{i_1}(z_1)\Phi_{i_2}(z_2) \cdots\Phi_{i_N}(z_N) \rangle
=-\displaystyle\sum_{j=1}^N\frac{(T^a)_{i_jk_j}}{(z-z_j)}
\langle\Phi_{k_1}(z_1)\Phi_{k_2}(z_2) \cdots\Phi_{k_N}(z_N) \rangle~,
\end{array}
\end{equation}
which are the ward identities corresponding to the conformal and gauge
symmetries.
One field theory possessing the above features is
the Wess-Zumino-Novikov-Witten (WZNW) model \cite{Novikov82,Feigin90,Gepner86}.
If dim($g$)=$d$, rank$(g)=l$, the Coulomb gas version of the
corresponding WZNW model is depicted by $(d-l)/2$
free $\beta$-$\gamma$ pairs and $l$ free real scalars $\vec{\phi}$ with a
boundary term (charge at infinity) \cite{Wakimoto,Feigin90,Gerasimov}.
The energy-momentum tensor is
\begin{equation}\label{4.2.55}
T(z) = - \sum_{\vec{\alpha}\in\Delta_+} :\beta_{\vec{\alpha}} \partial
\gamma_{\vec{\alpha}}:
- \frac{1}{2}(\partial\vec{\phi})^2 - \frac{i}{\nu}\vec{\rho} \cdot
\partial^2\vec{\phi}~,
\end{equation}
where $\nu=\sqrt{k+h^*}$, $h^*=\vec{\theta} \cdot (\vec{ \theta}+2\vec{\rho})/
\vec{\theta}^2$ is the dual Coxeter number, $\vec{\theta}$
is the highest root ($\vec{\theta}^2=2$) and
$2\vec{\rho}=\displaystyle\sum_{\vec{\alpha}\in\Delta_+}\vec{\alpha}$. \\
Thus, the OPEs are
\begin{equation}
\begin{array}{rcl}
& &\gamma_{\vec{\alpha}}(z)\beta_{\vec{\alpha}'}(w)=-\delta_{\vec{\alpha}\vec{
\alpha}'}\displaystyle\frac{1}{z-w}+ \cdots~,\\
& &\vec{ a}\cdot\vec{ \phi}(z)\vec{ b}\cdot\vec{ \phi}(w) =-\vec{ a}\cdot\vec{
b}
\log(z-w) + \cdots~,
\end{array}
\end{equation}
where $\vec{a}$ and $\vec{b}$ are arbitrary vectors in the roots space.\\
The central charge of this system indeed is
\begin{equation}
c=(d-l)+(l-\frac{12}{\nu^2}\vec{ \rho}^2)=\frac{kd}{k+h^*}~,
\end{equation}
where we have used the Freudenthal-de Vries formula $h^*=\displaystyle
\frac{12}{d}\vec{ \rho}^2$. The
vertex operators of primary fields are
\begin{equation}
V_{\vec{\lambda}}(z)=\exp\left((i/\nu)\vec{ \lambda}\cdot
\vec{\phi}(z)\right)
\end{equation}
with conformal weight
$$\Delta_d(\lambda)=\vec{ \lambda}\cdot(\vec{ \lambda}+2\vec{\rho})/(2\nu^2).$$
Their descendants with respect to the Kac-Moody algebra are
\begin{equation}
V_{\vec{\lambda},\vec{m}}(z)= P_{\vec{
m}}\left(\gamma(z)\right)V_{\vec{\lambda}}(z)~, \end{equation}
where $\vec{ m}$ is an element of the weight lattice. Unitary
modules correspond to background charges leading to rational central
terms, satisfying $k \in I -\{0\}$ and $\lambda_i \in I$. The Coulomb
gas duality shows up in that
$\Delta_d(V_{\vec{\lambda}}(z))=\Delta_d(V_{-2\vec{\rho}-\vec{ \lambda}}(z))$.
The screening currents associated with the simple roots acquire a compact
form
\begin{equation}
J_i(z)=R_i\left(\beta(z),~\gamma(z)\right)K_i(z)~,
\end{equation}
with $R_i$ a polynomial of degree one in the $\beta$'s (hence
$\Delta_d(R_i)=\Delta_d(\beta_i)=1$, since $\Delta_d(\gamma_i)=0$),
\begin{equation}
K_i(z)=\exp\left(-(i/\nu)\vec{\alpha}_i\cdot\vec{\phi}(z)\right)~,
\end{equation}
whose conformal weight $\Delta(K_i)_d=\vec{\alpha}_i\cdot
(\vec{ \alpha}_i-2\vec{\rho})/2\nu^2=0$.
The Kac-Moody currents can be classified into
$$\{x^+_i,~x^-_i,~k_i\}~,~~~~1\leq i\leq d~.$$
For the vertex operators $V_{\vec{\lambda},\vec{ m}}(z)$ and the screening
currents $J_i(z)$ we have the following braid relations
\begin{equation}\label{4.2.61}
\begin{array}{rcl}
& & V_{\vec{\lambda},\vec{ m}}(z) V_{\vec{\lambda}',\vec{ m}'}(w)=q^
{\vec{\lambda}\cdot\vec{\lambda}'/2}V_{\vec{\alpha}',\vec{
m}'}(w)V_{\vec{\alpha}
,\vec{ m}}(z)~,\\[2mm]
& & J_i(z)V_{\vec{ \lambda},\vec{ m}}(w)=q_i^{-\lambda_i/2}V_{\vec{\lambda}
,\vec{ m}}(w)J_i(z)~,\\[2mm]
& & J_i(z)J_j(w)=q_i^{a_{ij}/2}J_j(w)J_i(z)~,
\end{array}
\end{equation}
where $\lambda_i=2\vec{\lambda}\cdot\vec{\alpha}_i/\vec{\alpha}^2_i,~q=
e^{2\pi i/(k+h^*)}$, $q_i=q^{\vec{\alpha}^2_i/2}$ and
$a_{ij}=2\vec{\alpha}_i\cdot\vec{\alpha}_j/\vec{alpha}_i^2$. Since $h^*$ is
always an integer for the unitary representation
($k$ an integer) $q$ is a root of unity. More generally, $q$ is a
root of unity whenever the level $k$ is rational.
Let us define the screening charges by
\begin{equation}
Q_i = \int_{z_a}^{z_b} dzJ_i(z)~,
\end{equation}
so as to balance the total Coulomb charge of the correlation function for
maintaining transformation invariance, where $z_a$ and $z_b$ are arbitrary
insertion points $\{z_1,~z_2,~\cdots,~
z_N\}$. The most general setting for the correlation function in the
background charge formalism will be given by\\ \\
$\langle V_{\vec{\lambda}_1,\vec{ m}_1}(z_1) V_{\vec{\lambda}_2,\vec{
m}_2}(z_2)
\cdots V_{\vec{\lambda}_N,\vec{ m}_N}(z_N)\rangle$
\begin{equation}
\sim\langle V_{\vec{\lambda}_1,\vec{ m}_1}(z_1)Q_i(z_a,z_b)\cdots Q_j(z_c,z_d)
V_{\vec{ \lambda}_2,\vec{ m}_2}(z_2) \cdots Q_k(z_e,z_f)
V_{\vec{\lambda}_N,\vec{ m}_N}\rangle~,
\end{equation}
where $\displaystyle\sum^N_{i=1} \vec{m}_i = 0$, and the total Coulomb
charge on the r.h.s. plus the background charge vanishes.
Define the screening operators $X^-_i$ as
\begin{equation}
X^-_i V_{\vec{\lambda},\vec{ m}}(z) = \frac{1}{1-q^{-1}_i}\int_{G} dtJ_i(t)
V_{\vec{\lambda},\vec{ m}}(z)~\sim~Q_i(\infty,z)
V_{\vec{\lambda},\vec{ m}}(z),
\end{equation}
where the contour $G$ is the same as that we used in the minimal model case.
Then, the $N$-point conformal block are
\begin{equation}
\lim_{z_1\to\infty}\langle \tilde{V}_{\vec{\lambda}_1,\vec{m}_1}(z_1)
X^-_i\cdots
X^-_j V_{\vec{\lambda}_2,\vec{m}_2}(z_2)
\cdots X^-_k \cdots X^-_l V_{\vec{\lambda}_N,\vec{m}_N}(z_N)\rangle~,
\end{equation}
with $\displaystyle\sum^N_{i=1} \vec{m}_i=0$, and vanishing total Coulomb
charge.
Notice that in the above equation we have used the $SL(2,C)$ invariance.
This allows us to adopt the convention that one vertex will be taken in its
dual form with the definition for the dual of vertex operator
$$\tilde{V}_{\vec{\lambda},\vec{ m}}(z)=
V_{-\vec{\lambda}-2\vec{\rho},-\vec{ m}}(z)~.$$
We define the operators $k_i$ by
\begin{equation}
k_iV_{\vec{\lambda},\vec{m}}(z)=\exp\left(i\frac{\nu}{\vec{\alpha}^2_i}\oint
dz\partial \vec{\phi}_i(z)-1\right)V_{\vec{\lambda},\vec{m}}(z)~.
\end{equation}
$k_i$ and $X^-_i$ generate the Borel subalgebra of the quantum
group underlying the WZNW model, which turn out to be $U_q(g)$, where
$g$ is the zero mode of the Kac-Moody algebra $\hat{g}$
\begin{equation}
[k_i,k_j] = 0,~~~~~k_iX^-_jk_i^{-1}=q_i^{-1/2}X^-_j~.
\end{equation}
Define the adjoint action of the quantum group ad$X^-_i$ by
\begin{equation}
\left({\rm ad}X^-_i\right)X^-_j=\frac{1}{1-q_i^{-1}}\int_{G}dzX^-_iJ_j(z)~,
\end{equation}
yielding
\begin{equation}
({\rm ad}X^-_i)X^-_j = X^-_iX^-_j - q_i^{a_{ij}/2}X^-_jX^-_i~.
\end{equation}
The consistency of the Borel subalgebra generated by $k_i$ and $X^-_i$
is ensured by the Serre relations
\begin{equation}\label{4.2.73}
\left({\rm ad}X^-_i\right)^{1-a_{ij}}X^-_j = 0~,~~~~(i \neq j)~.
\end{equation}
Owing to the product rule
\begin{equation}
\left({\rm ad}X^-_i\right)\left(X^-_j X^-_k\right) = \left({\rm ad}X^-_i\right)
(X^-_j)X^-_k + q_i^{a_{ij}/2}X^-_j\left({\rm ad}X^-_i\right)(X^-_k)~,
\end{equation}
Equation (\ref{4.2.73}) yields the explicit form of the Serre relation
\begin{equation}
\sum^{1-a_{ij}}_{k=0}(-1)^k q_i^{-k(1-a_{ij}-k)}
\left\vert\left[\begin{array}{c}1-a_{ij}\\ k \end{array}\right]\right.
_{q_i^{-1}}(X^-_i)^kX^-_j(X^-_i)^{1-a_{ij}-k}= 0~,
\end{equation}
where $\displaystyle\left\vert\left[\begin{array}{c}
n\\
k\end{array}\right]_q\right.=\frac{|[n]_q!}{|[k]_q!|[n-k]_q}$.
The dual to $X^-_i,~X^+_i$, and $k_i$ generated the other Borel subalgebra
of the quantum group. We define $X^+_i$ by
\begin{equation}\label{4.2.74}
\nu^2K_i(\infty)X^+_iX^-_IV_{\vec{\lambda},\vec{m}}(z)
=(x^-_i)^{(0)}X^-_IV_{\vec{\lambda},\vec{m}}(z)-
X^-_I(x^-_i)^{(0)}V_{\vec{\lambda},\vec{m}}(z)~,
\end{equation}
where the short-hand $X^-_I=X^-_{i_1i_2\cdots i_n}=X^-_{i_1}
X^-_{i_2}\cdots X^-_{i_n}$ and $(x^-_i)^{(0)}$ is the zero mode of the
Kac-Moody currents
$x^-_i$. The acting of the zero mode of the Kac-Moody currents $(x^-_i)^{(0)}$
on vertex operators can be computed directly,\\ \\
$(x^-_i)^{(0)}(X^-_j)^nV_{\vec{\lambda},\vec{m}}(z)$
\begin{equation}
\begin{array}{rcl}
&
&=[(x^-_i)^{(0)},(X^-_j)^n]V_{\vec{\lambda},\vec{m}}(z)+(X^-_j)^n(x^-_i)^{(0)}
V_{\vec{\lambda},\vec{m}}(z)\\[2mm]
& &=\nu^2\delta_{ij}K_i(\infty)\displaystyle\sum^{n-1}_{k=0}
q^{n-1-k}_i(X^-_i)^{n-1-k}\frac{1-k_i^{-1}}{1-q^{-1}_i}
(X^-_i)^kV_{\vec{\lambda},\vec{m}}(z)+(X^-_j)^n(x^-_i)^{(0)}V_{\vec{\lambda}
,\vec{m}}(z)~.
\end{array}
\end{equation}
Then we have
\begin{equation}
X^+_iV_{\vec{\lambda},\vec{m}}(z)=0~,
\end{equation}
and
\begin{equation}
\begin{array}{rcl}
& &k_iX^+_jk_i^{-1}=q_i^{1/2}X^+_j~,\\[2mm]
&
&X^+_iX^-_j-q^{a_{ij}/2}_iX^-_jX^+_i=\delta_{ij}\displaystyle\frac{1-k_i^{-2}}
{1-q^{-1}_i}~.
\end{array}
\end{equation}
The adjoint action of $X^+_i$ on $X^+_j$ is also given by
\begin{equation}
\left({\rm ad}X^+_i\right)X^+_j = X^+_iX^+_j - q_i^{a_{ij}/2}X^+_jX^+_i~.
\end{equation}
Consequently, the corresponding Serre relations are
\begin{equation}
\left({\rm ad}X^+_i\right)^{1-a_{ij}}X^+_j=0~.
\end{equation}
The explicit form of the Serre relations are
\begin{equation}
\sum^{1-a_{ij}}_{k=0}(-1)^k q^{-k(1-a_{ij}-k)}_i
\left\vert\left[\begin{array}{c}
1-a_{ij}\\
k\end{array}\right]_{q^{-1}_i}\right.
(X^+_i)^kX^+_j(X^+_i)^{1-a_{ij}-k}=0~.
\end{equation}
Simple contour deformations are sufficient for finding the comultiplication,
\begin{equation}
\begin{array}{rcl}
& &\Delta(k_i)=k_i \otimes k_i~,\\[2mm]
& &\Delta(X^\pm_i)=X^-_i \otimes 1 +k_i^{-1} \otimes X^-_i~,\\[2mm]
& &\Delta(X^+_i)=X^+_i \otimes 1 + k_i^{-1}\otimes X^+_i~.
\end{array}
\end{equation}
The last comultiplication, for $X^+_i$, arises from an integration likewise in
Fig.(21),
however, due to th factor $K_i(\infty)$ in Eq.(\ref{4.2.74}), we have the
braiding phase that makes the comultiplication non-commutative.
\setlength{\unitlength}{5pt}
\thicklines
\begin{picture}(50,13)(0,4)
\put(10,10){\oval(15,5)}
\put(7,10){$\bullet$}
\put(8.5,9.4){$z_1$}
\put(12,10){$\bullet$}
\put(13.5,9.4){$z_2$}
\put(8,5){$\Delta G$}
\put(21.5,9.5){$=$}
\put(35,10){\oval(8,5)}
\put(34.5,10){$\bullet$}
\put(36,9.4){$z_1$}
\put(43.5,10){$\bullet$}
\put(45,9.4){$z_2$}
\put(34,5){$G_1$}
\put(69,10){\oval(8,5)}
\put(53,9.5){$+$}
\put(59.5,10){$\bullet$}
\put(61,9.4){$z_1$}
\put(68.5,10){$\bullet$}
\put(70,9.4){$z_2$}
\put(69,5){$G_2$}
\end{picture}
\hspace{7.5em}Fig.21. Comultiplication operation of $X_i^+$.
In this basis the counit $\epsilon$ and antipode $S$ are,
\begin{equation}
\begin{array}{l}
\epsilon(k_i)=1~,~~~~ \epsilon(X^-_i)=0=\epsilon(X^+_i)~,\\
S(k_i)=k_i^{-1}~,~~~S(X^-_i)=-X^-_ik_i^{-1}~,~~~S(X^+_i) = -k_iX^+_i~.
\end{array}
\end{equation}
Now we discuss the contour representations of the quantum group $U_q(g)$.
Let us first review the simplest case of $SU_q(2)$, from which the general
case follows, mainly because any Lie algebra can be viewed as a
superposition of the $SU(2)$'s associates with the various simple
roots. The conventional notation $\lambda=2j$ with $j$ the spin
of the representation is adopted. The braidings among vertex operators
$V_j={\rm
exp} (ij\phi/\nu)$ of spin $j$ and the screening current $J=\beta{\rm exp}
(-i\phi/\nu)$ follow from relation (\ref{4.2.61}),
\begin{equation}
\begin{array}{rcl}
& &V_j(z)V_{j'}(w)=q^{jj'}V_{j}(w)V_j(z)~,\\
& &J(z)V_j(w)=q^{-j}V_j(w)J(z)~,\\
& &J(z)J(w)=qJ(w)J(z)~.
\end{array}
\end{equation}
The screened vertex operators $e_n^j(z)$ give the representations of the
quantum group $SU_q(2)$
\begin{equation}
\begin{array}{l}
e^j_n(z)=\left(X^-\right)^nV_j(z)~,\\
X^-e^j_n(z)=e^j_{n+1}(z)~,\\
X^+e^j_n(z)=|[n]_{q^{-1}}|[2j-n+1]_qe^j_{n-1}(z)~,\\
ke^j_n=\exp(j-n)e^j_n~.
\end{array}
\end{equation}
By making use of path-ordering, these operators can be rewritten
in terms of the actions of integrals
$P=\displaystyle\int_\infty^zdtJ(t)$ on $V_j(z)$,
\begin{equation}
e_n^j = (X^-)^nV_j(z) =|[2j]_q|[2j-1]_q \cdots|[2j-n+1]_q P^nV_j(z)~.
\end{equation}
Clearly, if $n>2j$, $e_n^j=0$, and $\dim\{(e_n^j)\}=2j+1$.
It is reasonable to interpret $e_n^j$ as a $q$-multiplet of spin
$j$.
The time-ordering of the screenings is
\begin{equation}
P^n=|[n]_{q^{-1}}!T(P^n)=|[n]_{q^{-1}}!\int_\infty^zdt_1\int_\infty^{t_1}dt_2
\cdots\int_\infty^{t_{n-1}}dt_nJ(t_n)\cdots J(t_1)~.
\end{equation}
Similarly, for an arbitrary simple group $X^-_{i_1}\cdots
X^-_{i_n} V_{\vec{\lambda},\vec{m}}(z)$ can be decomposed into products of
integrals
like
\begin{equation}
P_i = \int^z_\infty dtJ_i(t)~,
\end{equation}
by means of the recursion formula \\ \\
$X^-_iP_{i_1} \cdots P_{i_n} V_{\vec{\lambda},\vec{m}}(z)$
\begin{equation}
=\frac{1}{1-q^{-1}_i}\left(P_iP_{i_1} \cdots P_{i_n}
V_{\vec{\lambda},\vec{m}}(z)
-q^{-\lambda_i+(a_{ii_1}+
\cdots +a_{ii_n})/2}_i P_{i_1}\cdots
P_{i_n}P_i\right)V_{\vec{\lambda},\vec{m}}(z)~.
\end{equation}
Furthermore, the products of screening $P_i$ must be time-ordered.
Consider, for example, a product of two $P_i$. It is simple to
see that
\begin{equation}
P_{i_1}P_{i_2}V_{\vec{\lambda},\vec{m}}(z) = \left(T\left(P_{i_1}P_{i_2}\right)
-q^{a_{i_1i_2}/2}_{i_1}T\left(P_{i_2}P_{i_1}\right)\right)
V_{\vec{\lambda},\vec{m}}(z)~,
\end{equation}
and the following recursion relation in general holds
\begin{equation}
P_iT(P_{i_1}\cdots P_{i_n})V_{\vec{\lambda},\vec{m}}(z)
=\sum^n_{k=0}q_i^{(a_{ii_1}+\cdots+a_{ii_k})/2}
T(P_{i_1}P_{i_2}\cdots P_{i_k}P_iP_{i_{k+1}}\cdots
P_{i_n}V_{\vec{\lambda},\vec{m}}(z)~,
\end{equation}
where
\begin{equation}
T(P_{i_1}\cdots P_{i_n}) = \int^z_\infty dt_n \int^{t_n}_\infty
dt_{n-1}\cdots \int^{t_2}_\infty
dt_1J_{i_1}(t_1)J_{i_2}(t_2)\cdots J_{i_n}(t_n)~.
\end{equation}
The resultant contour representation of the quantum group is
consistent, i.e, it fulfill the Serre relation.
\section{Molecular spectrum}
The Schr\"{o}dinger equation for a diatomic molecule is
\begin{equation}
\frac{1}{m}\sum_i\left(\frac{\partial^2\Psi}{\partial x_i^2}
+\frac{\partial^2\Psi}{\partial y_i^2}+\frac{\partial^2\Psi}{\partial z_i^2}
\right)+
\sum_k\frac{1}{M_k}\left(\frac{\partial^2\Psi}{\partial x_k^2}+
\frac{\partial^2\Psi}{\partial y_k^2}+\frac{\partial^2\Psi}{\partial z_k^2}
\right)+\frac{8\pi^2}{h^2}\left(E-V\right)\Psi=0~,
\end{equation}
where $x_i,y_i,z_i$ are coordinates for electrons with identical
masses $m$, while
the $x_k,y_k,z_k$ are coordinates for the nuclei with masses $M_k$.
According to Born and Oppenheimer approximation \cite{Born},,
the wave function $\Psi$ can be separated into
\begin{equation}\label{6.4}
\Psi=\psi_e\left(\cdots,x_i,y_i,z_i,\cdots
\right)\psi_{\rm vib-rot}\left(\cdots,x_k,y_k,z_k,\cdots\right)~,
\end{equation}
where the $\psi_e,~\psi_{\rm vib-rot}$ are the solutions of the following
equations,
\begin{equation}\label{6.5}
\begin{array}{l}
\displaystyle \sum_i\left(\frac{\partial^2\psi_e}{\partial x_i^2}+
\frac{\partial^2\psi_e}{\partial y_i^2}+\frac{\partial^2\psi_e}{\partial z_i^2}
\right)+\frac{8\pi^2m}{h^2}\left(E_{e}-V_e\right)\psi_e=0~,\\[2mm]
\displaystyle \sum_k \frac{1}{M_k}\left(\frac{\partial^2\psi_{\rm
vib-rot}}{\partial x_k^2}+
\frac{\partial^2\psi_{\rm vib-rot}}{\partial y_k^2}+\frac{\partial^2\psi_{\rm
vib-rot}}
{\partial z_k^2}
\right)+\frac{8\pi^2}{h^2}\left(E-E_{e}-V_n\right)\psi_{\rm vib-rot}=0~.
\end{array}
\end{equation}
The first equation is the Schr\"{o}dinger equation of electrons
moving in the field of fixed nuclei, represented by an effective potential
$V_e$. The second one describes the motion in the
effective potential $E_{e}+V_n$, with $V_n=\displaystyle\frac{z_1z_2e^2}{r}$,
the Coulomb potential between the two nuclei of electric charge $z_1e$
and $z_2e$ in distance $r$. It can be cast into the
following form
\begin{equation}
\left(-\frac{h^2}{8\pi^2 M_1}\bigtriangledown _1^2-
\frac{h^2}{8\pi^2 M_2}\bigtriangledown _2^2+V_n(r)\right)\psi_{\rm vib-rot}
=E_t\psi_{\rm vib-rot}~,
\end{equation}
where $E_t=E-E_e$. In the center-of-mass frame the equation reads
\begin{equation}
\left(-\frac{h^2}{8\pi^2 M}\bigtriangledown^2+V_n(r)\right)
\psi_{\rm vib-rot}=E_t\psi_{\rm vib-rot}~,
\end{equation}
where $\displaystyle M=\frac{M_1M_2}{M_1+M_2}$ is the reduced mass. As the
tangent and radial variables can be separated, we have
\begin{equation}
\psi_{\rm vib-rot}=R(r)\psi_{\rm rot}(\theta,\phi)~, ~~R(r)={1\over r}
\psi_{\rm vib}(r)~,
\end{equation}
and
\begin{equation}\label{6.8}
\Psi=\psi_e\cdot {1\over r}\psi_{\rm vib}\psi_{\rm rot}~.
\end{equation}
The Hamiltonian of the diatomic molecule are composed of three parts
for electronic transition, vibration and rotation, i.e.,
\begin{equation}\label{6.9}
H=H_e+H_{\rm vib}+H_{\rm rot}~.
\end{equation}
The phenomenological description of the vibrational and rotational
energy is given by the following formulas \cite{Herzberg}--\cite{Mizushima}
\begin{equation}\label{6.22}
\begin{array}{l}
E_{\rm vib}=\displaystyle hc\omega_e(v+{1\over 2})-hc\omega_ex_e(v+{1\over
2})^{2}
+hc\omega_ey_e(v+{1\over 2})^3+\cdots~,\\[2mm]
E_{\rm rot}=BhcJ(J+1)-DhcJ^2(J+1)^2+HhcJ^3(J+1)^3+\cdots~.
\end{array}
\end{equation}
When the coefficients are selected appropriately, the spectra given by the
above formulas may fit with experimental results very well.
The interaction between vibration and rotation can be taken into account by
the explicit $v$ dependence of the coefficients $B$, $D$, $\cdots$, in the
second formula of Eq.(\ref{6.22}).
To lowest order, the set of vibration-rotational constants can be
written as
\begin{equation}\begin{array}{rcl}
B_v&=&B-\alpha_e(v+\displaystyle\frac{1}{2})+\cdots~,\\[2mm]
D_v&=&D+\beta_e(v+\displaystyle\frac{1}{2})+\cdots~,
\end{array}\end{equation}
where $\alpha_e$ and $\beta_e$ are constants much smaller than $B$ and $D$
respectively.
Thus, we have the energy levels of the vibrating and rotating diatomic
molecule
\begin{equation}\label{6.6429}
\begin{array}{rcl}
E_{\rm vib-rot}(v,J)&=&\displaystyle hc\omega_e\left(v+{1\over 2}\right)
-hc\omega_ex_e\left(v+{1\over 2}
\right)^{2}+hc\omega_ey_e\left(v+{1\over 2}\right)^3\\[2mm]
&&\displaystyle-hc\alpha_e(v+\frac{1}{2})J(J+1)
-hc\beta_e(v+\frac{1}{2})J^2(J+1)^2\\[2mm]
&&+hcB_eJ(J+1)-hcD_eJ^2(J+1)^2+\cdots~.
\end{array}\end{equation}
The complete phenomenological description for vibrational and rotational
structure of diatomic molecule is given by the Dunham expansion
\cite{Dunham29,Dunham32}
\begin{equation}
E_{\rm vib-rot}(v,J)=hc\sum_{ij}Y_{ij}\left(v+\frac{1}{2}\right)^i
\left(J(J+1)\right)^j~,
\end{equation}
where $Y_{ij}$ are the coefficients of the Dunham expansion. A successful
theory for vibrating and rotating spectra of diatomic molecules
must recover leading terms of the Dunham expansion.
\subsection{Vibrating diatomic molecular spectrum}
The Hamiltonian for a quantum $q$-oscillator system \cite{ZC91a,ZC91b} is
\begin{equation}\label{6.23}
H_{q-{\rm vib}}={1\over 2}\left(a_q^\dagger a_q+
a_q a_q^\dagger \right)hc\nu_{vib}~,
\end{equation}
where $a_q$, $a_q^\dagger$ are annihilation and creation
operators for the deformed system. These operators are related with
the operators $a,a^\dagger$ of the harmonic oscillator system by
\begin{equation}\label{6.6432}
a_q=\sqrt{\frac{[N +1+b\gamma]_q}{N +1}}a~,~~~~
a_q^\dagger=a^\dagger\sqrt{\frac{[N+1+b\gamma]_q}{N+1}}~,
\end{equation}
where $N =a^\dagger a$. Making use of the basic commutators
\begin{equation}\label{6.26}
[a,a^\dagger]=1,~~~[a,a]=[a^\dagger,a^\dagger]=0,
\end{equation}
we have the following commutation relations \cite{Macfarlane}--\cite{Song90}
\begin{equation}\label{6.37}
\begin{array}{rcl}
\displaystyle \left[a_q,a_q^\dagger\right]&=&\left[N+1+b\gamma\right]_q
-\left[N+b\gamma\right]_q~,\\
\displaystyle \left[N, a_q\right]&=&-a_q~,~~~
\displaystyle \left[N, a_q^\dagger\right]=a_q^\dagger~.
\end{array}\end{equation}
The Hopf algebra structure can be defined by
\begin{equation}\label{6.37a}
\begin{array}{l}
\Delta(N')=N'\otimes 1+1\otimes N'-i\displaystyle\frac{\alpha}{\gamma}
1\otimes 1~,\\
\Delta(a_q)=\left(a_q\otimes q^{N'/2}+iq^{-N'/2}\otimes
a_q\right)e^{-i\alpha/2}~,\\
\Delta(a_q^\dagger)=\left(a_q^\dagger\otimes q^{N'/2}+
iq^{-N'/2}\otimes a_q^\dagger\right)e^{-i\alpha/2}~,\\
\epsilon(N')=\displaystyle i\frac{\alpha}{\gamma}~,~~~~
\epsilon(a_q)=0=\epsilon(a_q^\dagger)~,\\
S(N')=-N'+i\displaystyle\frac{2\alpha}{\gamma}\cdot 1~,\\
S(a_q)=-q^{-1/2}a_q~,~~~~~S(a_q^\dagger)=-q^{1/2}a_q^\dagger~,
\end{array}
\end{equation}
where $N'=N+\displaystyle\frac{1}{2}+b\gamma,~\alpha=2k\pi+\frac{\pi}{2},
{}~k\in Z$.\\
Eqs.(\ref{6.37}) and (\ref{6.37a}) constitute the quantum Weyl-Heisenberg
group $H_q(4)$.
Use is made of Eq.({\ref{6.6432}) to cast the Hamiltonian for the
$q$-oscillator system into
\begin{equation}\label{6.e316}
H_{q-{\rm vib}}={1\over 2}\left([N+b\gamma ]_q+[N+1+b\gamma
]_q\right)hc\nu_{vib}~.
\end{equation}
The representations of quantum group $H_q(4)$ are constructed by
\begin{equation}\label{6.29}
\mid{v}\rangle\rangle=\left([v+b\gamma ]_q!\right)^{-1/2}\left(a_q^\dagger
\right)^v\vert 0\rangle~.
\end{equation}
The actions of the operators $a_q^\dagger$, $a_q$ on the Fock states yield
\begin{equation}\label{6.30}
\begin{array}{l}
\displaystyle
a_q^\dagger\vert {v}\rangle\rangle=\sqrt{[v+1+b\gamma ]_q}\vert v+1\rangle
\rangle~,\\
\displaystyle a_q\vert {v}\rangle\rangle=\sqrt{[v+b\gamma ]_q}\vert v-1
\rangle\rangle~,\\
\displaystyle a_q\vert {0}\rangle=0~,
\end{array}\end{equation}
and the energy levels of the system are
\begin{equation}\label{6.31}
\begin{array}{rcl}
E_{q-{\rm vib}}(v)&=&\displaystyle{1\over 2}\left([v+1+b\gamma ]_q+[v+b\gamma
]_q\right)h
c\nu_{vib}\\[2mm]
&=&\displaystyle\frac{hc\nu_{\rm
vib}}{2\sinh\frac{\gamma}{2}}
\sinh\left(\gamma(v+\frac{1}{2}+b\gamma)\right)\\[2mm]
&=&\displaystyle hc\nu_{\rm vib}\left(\frac{\sinh(\gamma c)}
{2\sinh\left(\frac{\gamma}{2}\right)}+\cosh(\gamma
c)\left(
v+\frac{1}{2}\right)\right.\\[2mm]
& &~~~~~\displaystyle\left.+\frac{\gamma\sinh(\gamma c)}{2}
\left(v+\frac{1}{2}\right)^2+\frac{\gamma^2\cosh(\gamma
c)}{6}
\left(v+\frac{1}{2}\right)^3+\cdots\right)~,
\end{array}
\end{equation}
where $c\equiv b\gamma$.
It is easy to see that the above equation gives the coefficients
$\omega_e,~\omega_ex_e,~\omega_ey_e,~\cdots$ of Eq.({\ref{6.22}).
For generic $q$, the representation for the quantum group
$H_q(4)$ is isomorphic to that for the Lie group $H(4)$
\begin{equation}\label{6.32}
\begin{array}{rcl}
\vert v\rangle\rangle&=&\left([v+b\gamma ]_q !\right)^{-1/2}
\left(a_q^\dagger\right)^v\vert 0\rangle\\
&=&\left(v!\right)^{-1/2}\left(a^\dagger\right)^v
\vert 0\rangle\\
&=&\vert v\rangle~.
\end{array}\end{equation}
where $\vert v\rangle$ are the Fock states for the harmonic oscillator system.
Therefore, the representations of $H_q(4)$
in space coordinates can be exactly expressed by Hermite polynomials, i.e.,
\begin{equation}\label{6.33}
\tilde{\psi}_v(x)= N_v H_v(X)e^{-X^2/2}~,
\end{equation}
where $X=\beta x$, $x$ is the change of
internuclear distance away from equilibrium position,
$\beta=\left((2\pi)^2mc\nu_{\rm vib}/h\right)^{1/2}$ and $N_v=\left(\sqrt{\pi}
2^vv!\right)^{-1/2}$.
If there is a dipole moment for the nuclei in equilibrium position, as
it should be for the molecule of unlike atoms, this dipole
moment changes upon varying the internuclear distance. As a first
order approximation, the dipole moment is assumed to vary linearly
with the internuclear distance, i.e., $M=M_0+M_1x$,
where $M_0$ is the dipole moment at equilibrium, $M_1$ is the rate of
change of the dipole moment with the
internuclear distance. Therefore, the transition matrix elements are
\begin{equation}\label{6.35}
R^{v^\prime v^{\prime\prime}}_q=\displaystyle
\int\tilde{\psi}^\ast_{v^\prime} M
\tilde{\psi}_{v^{\prime\prime}} dx =
\displaystyle M_0\delta_{v^\prime v^{\prime\prime}}+M_1\frac{N_{v^\prime}
N_{v^{\prime\prime}}}{\beta}\int X H_{v^\prime}(X) H_{v^{\prime\prime}}
(X) e^{-X^2}dx~.
\end{equation}
By the recursion relation of the Hermite polynomial, the second term
in the above equation vanishes unless
$v^\prime=v^{\prime\prime}\pm 1$. The selection rule
for the $q$-oscillator system is $\Delta v=\pm 1$. Therefore the infrared
spectrum of the $q$-oscillator system is
\begin{equation}
\label{6.39}
\nu=\frac{1}{2}\left([v+2+b\gamma ]_q-[v+b\gamma ]_q\right)\nu_{vib}~.
\end{equation}
An external electric field ${\bf F}$ induces
a dipole moment ${\bf p}$ in the diatomic system. Its magnitude
is proportional to that of the field,
\begin{equation}\label{6.40}
\vert{\bf p}\vert=\alpha\vert {\bf F}\vert~,
\end{equation}
where $\alpha$ is the polarizability.
The transition matrix elements corresponding to the induced dipole moment
are
\begin{equation}\label{6.41}
\left|{\bf p}\right|^{v^\prime v^{\prime\prime}}=\int \tilde{\Psi}_{v^\prime}
^\ast {\bf p}\tilde{\Psi}_{v^{\prime\prime}}dx~,
\end{equation}
where $\tilde{\Psi}_{v^\prime}$ and $\tilde{\Psi}_{v^{\prime\prime}}$
are the (time dependent)
wave functions of the $q$-oscillator system at states $v^\prime$ and
$v^{\prime\prime}$ \ respectively. The evolution factors for
$\tilde{\Psi}_{v^\prime}^\ast$, $\tilde{\Psi}_{v^{\prime\prime}}$ and
${\bf p}$ should be $e^{2\pi i(E_{q-{\rm vib}}(v^\prime)/h)t}$,
$e^{-2\pi i(E_{q-{\rm vib}}(v^{\prime\prime})/h)t}$ and $e^{2\pi ic
\nu_{\rm ext}t}$,
where the $c\nu_{\rm ext}$ is the frequency of the external electric field.
$\left|{\bf p}\right|^{v^\prime v^{\prime\prime}}$ evolves by
frequency $c\nu_{\rm ext}+(E_{q-{\rm vib}}(v^\prime)-E_{q-{\rm vib}}(
v^{\prime\prime}))/h$ with the amplitude
\begin{equation}\label{6.42}
\left|{\bf p}\right|^{v^\prime v^{\prime\prime}}=
\vert {\bf F}\vert \int \tilde{\psi}_{v^\prime}^\ast \alpha
\tilde{\psi}_{v^{\prime\prime}}dx~.
\end{equation}
To the lowest order assuming a linear variation of $\alpha$
with the displacement $x$ from the equilibrium position, i.e.,
\begin{equation}\label{6.43}
\alpha=\alpha_{0v}+\alpha^1_vx~,
\end{equation}
we have
\begin{equation}\label{6.e23}
\left|{\bf p}\right|^{v^\prime v^{\prime\prime}}=
\vert {\bf F}\vert\alpha_{0v}\int
\tilde{\psi}^\ast_{v^\prime}\tilde{\psi}_{v^{\prime\prime}}dx+
\vert {\bf F}\vert\alpha_{v}^1\int x
\tilde{\psi}^\ast_{v^\prime}\tilde{\psi}_{v^{\prime\prime}}dx~.
\end{equation}
Because of the orthogonality of the wave functions of the $q$-oscillator
system, the first term in the above equation is zero unless
$v^\prime=v^{\prime\prime}$, which gives the Reighley scattering. The
integration in the second
term vanishes unless $v^\prime=v^{\prime\prime}\pm 1$. The selection
rule for vibrational Raman spectrum is $\Delta v=\pm 1$.
We then obtain the vibrational Raman
spectrum from the energy levels given in Eq.(\ref{6.e316}),
\begin{equation}\label{6.46}
\begin{array}{rcl}
\nu&=&{\nu^\prime}\pm \left(E_{q-{\rm vib}}(v^\prime+1)-E_{q-{\rm vib}}
(v)\right)\\[3mm]
&=&\displaystyle{\nu^\prime}\pm \frac{1}{2}\left([v+2+b\gamma ]_q-[v+b\gamma
]_q\right)\nu_{vib}~,
\end{array}
\end{equation}
where $\nu^\prime$ is the wave number of the incident photon.
The electronic transitions are involved in the
visible and ultraviolet spectral regions. Denote the vibrational spectrum of
the
electronic state by $E_{\rm e-vib}$, leading to
\begin{equation}\label{6.56x}
E_{\rm e-vib}=E_0+E_{\rm q-vib}~.
\end{equation}
Then, the vibrational spectrum of electronic transitions is
\begin{equation}\label{6.47}
\begin{array}{rcl}
\nu&=&\left(E^\prime_{\rm e-vib}-E^{\prime\prime}_{\rm e-vib}\right)/hc\\
&=&(E_0^\prime-E_0^{\prime\prime})/hc
+(G^\prime_{q^\prime}-G^{\prime\prime}_{q^{\prime\prime}})
=\nu_e+\nu_{v}~,
\end{array}
\end{equation}
where $\nu_e$ is a certain constant depending on the transition of
electronic states, $G_q\equiv E_{q-vib}/hc$. By Eq.(\ref{6.e316}) we rewrite
Eq.(\ref{6.47}) into the following:
\begin{equation}\label{6.48}
\begin{array}{rcl}
\nu&=&\displaystyle\nu_e+\frac{1}{2}\left([v^\prime+1+b^\prime\gamma^\prime
]_{q^\prime}+[v^\prime+b^\prime\gamma^\prime ]_{q^\prime}\right)
\nu_{vib}^\prime\\[2mm]
& &-\displaystyle\frac{1}{2}\left(
[v^{\prime\prime}+1+b^{\prime\prime}\gamma^{\prime\prime} ]_{q^{\prime\prime}}
+[v^{\prime\prime}+b^{\prime\prime}\gamma^{\prime\prime}
]_{q^{\prime\prime}}\right)\nu_{vib}^{\prime\prime}~.
\end{array}
\end{equation}
An investigation of the selection rules shows that for electronic transitions
there is no strict selection rule for the vibrational quantum number $v$.
In principle, each vibrational state of the upper electronic state can be
combined with each vibrational state of lower electronic
state, i.e., there is no restriction to the quantum numbers $v^\prime$ and
$v^{\prime\prime}$ in Eq.(\ref{6.48}), and therefore Eq.(\ref{6.48}) gives very
complicated
spectral structures.
If the quantum number $v^\prime$ in Eq.(\ref{6.48}) is fixed, then the
$v^{\prime\prime}$ progression
is formed. In the $v^{\prime\prime}$ progression
the upper vibrational state is fixed while the lower vibrational state
is different. Then Eq.(\ref{6.48}) is rewritten in the following form
\begin{equation}\label{6.49}
\nu={\nu_0^{\prime\prime}}-{1\over 2}\left([v^{\prime\prime}+1+b^{\prime\prime}
\gamma^{\prime\prime} ]_{q^{\prime\prime}}
+[v^{\prime\prime}+b^{\prime\prime}\gamma^{\prime\prime}
]_{q^{\prime\prime}}\right){\nu_{vib}^{\prime\prime}}~,
\end{equation}
where the quantum number ${\nu_0^{\prime\prime}}$ is $\nu_e$
plus the fixed vibrational
spectrum in the upper electronic state, and therefore is a constant.
If the quantum number $v^{\prime\prime}$ is chosen to be a constant, the
$v^\prime$ progression is formed in which different vibrational state
in an upper electronic state combine with the vibrational state of
lower electronic state. It is expressed by the following formula
\begin{equation}
\label{6.50}
\nu={\nu_0^{\prime}}+{1\over 2}\left([v^{\prime}+1+b^{\prime}\gamma^{\prime}
]_{q^{\prime}}
+[v^{\prime}+b^{\prime}\gamma^{\prime}
]_{q^{\prime}}\right){\nu_{vib}^\prime}~,
\end{equation}
where $\nu_0^\prime$ is $\nu_e$ minus the fixed vibrational spectrum of
the lower electronic state, and therefore is a constant.
\subsection{Rotating diatomic molecular spectrum}
The Hamiltonian for the $q$-rotator system \cite{ZC91b},\cite{ZC91bb}--
\cite{Bonatsos90} is
\begin{equation}\label{6.1a}
H_{q-{\rm rot}}=\frac{h^2}{8\pi^2I}C_{I,q}~,
\end{equation}
where
\begin{equation}\label{6.2a}
C_{I,q}=J^\prime_- J^\prime_++[J_3^\prime]_q[J_3^\prime+1]_q~,
\end{equation}
is the Casimir operator for the quantum group $SU_q(2)$
\begin{equation}
\begin{array}{l}
[J_3',J_\pm']=\pm J_\pm'~,\\[1mm]
[J_+',J_-']=[2J_3']~,\\[1mm]
\Delta(J_3')=J_3'\otimes 1+1\otimes J_3'~,\\
\Delta(J_\pm')=J_\pm'\otimes q^{-J_3'}+q^{J_3'}\otimes J_\pm'~,\\
\epsilon(J_3')=0=\epsilon(J_\pm')~,\\
S(J_3')=-J_3'~,~~~~~S(J_\pm')=-q^{\pm 1}J_\pm'~.
\end{array}
\end{equation}
It is well-known that the generators of the quantum group $SU_q(2)$ are
related with the generators of the Lie group $SU(2)$ by
\begin{equation}\label{6.8a}
\begin{array}{l}
J_+^\prime=\displaystyle
\sqrt{\frac{[J_3+J_0]_q[J_3-1-J_0]_q}{(J_3+J_0)
(J_3-1-J_0)}}J_+~,\\[3mm]
J_-^\prime=\displaystyle J_-\sqrt{\frac{[J_3+J_0]_q[J_3-1-J_0]_q}{(J_3+J_0)
(J_3-1-J_0)}}~,\\[3mm]
J_3^\prime=J_3~.
\end{array}
\end{equation}
For generic $q$, the representations (in coordinates space) of
$SU_q(2)$ can be chosen to be the spherical harmonics , i.e.,
\begin{equation}\label{6.9a}
\tilde{\psi}_{JM}({\bf x})=Y_{JM}(\theta, \phi)~.
\end{equation}
The actions of the generators of $SU_q(2)$ yield
\begin{equation}\label{6.10a}
\begin{array}{l}
J^\prime_\pm\tilde{\psi}_{JM}({\bf x})=\sqrt{[J\mp M]_q
[J\pm M+1]_q}\tilde{\psi}_{J,M\pm 1}({\bf x})~,\\[2mm]
J^\prime_3\tilde{\psi}_{JM}({\bf x})=M\tilde{\psi}_{JM}({\bf x})~.
\end{array}
\end{equation}
The action of the Casimir operator $C_{I,q}$ yields
\begin{equation}\label{6.11a} C_{I,q}\tilde{\psi}_{JM}({\bf
x})=[J+1]_q[J]_q\tilde{\psi}_{JM}({\bf x})~.
\end{equation}
Therefore, the
eigenvalues of the Hamiltonian for the $q$-rotator system are
\begin{equation}\label{6.ppa}
\begin{array}{rcl}
E_{q-{\rm rot}}&=&\displaystyle\frac{h^2}{8\pi^2I}[J]_q[J+1]_q\\[2mm]
&=&\displaystyle\frac{h^2}{8\pi^2I}\left(\left(1-\frac{1}{6}
\gamma^2+\frac{7}{360}\gamma^4\right)J(J+1)+
\gamma^2\left(\frac{1}{3}-\frac{7}{90}\gamma^2\right)\left(
J(J+1)\right)^2\right.\\[2mm]
& &\displaystyle\left. +\frac{2\gamma^4}{45}\left(J(J+1)
\right)^3+\cdots\right)~.
\end{array}
\end{equation}
It is easy to see that Eq.(\ref{6.ppa}) gives the coefficients
$B_0,~D_0,~H_0,~\cdots$ of Eq.(\ref{6.22}).
There is internal dipole moment $M_0$ in the diatomic molecules
system, as it is always for molecules consisting of unlike atoms
with spatial components,
\begin{equation}\label{6.12a}
\begin{array}{l}
M_{0x}=M_0\sin\theta\cos\phi~,\\[2mm]
M_{0y}=M_0\sin\theta\sin\phi~,\\[2mm]
M_{0z}=M_0\cos\theta~.
\end{array}\end{equation}
The dipole transition matrix elements are
\begin{equation}\label{6.13a}
\begin{array}{l}
R_x^{J^\prime M^\prime J^{\prime\prime} M^{\prime\prime}}
=\displaystyle M_0\int \tilde{\psi}_{J^\prime M^{\prime}}^\ast
\sin\theta\cos\phi\tilde{\psi}_{J^{\prime\prime}M^{\prime\prime}}
d\tau~,\\[2mm]
R_y^{J^\prime M^\prime J^{\prime\prime} M^{\prime\prime}}
=\displaystyle M_0\int \tilde{\psi}_{J^\prime M^{\prime}}^\ast
\sin\theta\sin\phi\tilde{\psi}_{J^{\prime\prime}
M^{\prime\prime}}d\tau~,\\[2mm]
R_z^{J^\prime M^\prime J^{\prime\prime} M^{\prime\prime}}
=\displaystyle M_0\int \tilde{\psi}_{J^\prime M^{\prime}}^\ast
\cos\theta\tilde{\psi}_{J^{\prime\prime}M^{\prime\prime}}d\tau~,
\end{array}
\end{equation}
where $d\tau=\sin\theta d\theta d\phi$.
Applying the recursion relation of the spherical harmonics,
we can cast $R_z^{J^\prime M^\prime J^{\prime\prime} M^{\prime\prime}}$ into
\begin{equation}
\begin{array}{rcl}
R_z^{J^\prime J^\prime J^{\prime\prime} M^{\prime\prime}}
&=&M_0\left(a_{J^{\prime\prime},M^{\prime\prime}}\displaystyle
\int Y_{J^\prime M^\prime}^\ast Y_{J^{\prime\prime}+1, M^{\prime\prime}}
\sin\theta d\theta d\phi\right.\\[4mm]
& &\left.+a_{J^{\prime\prime}-1,M^{\prime\prime}}\displaystyle
\int Y_{J^\prime, M^\prime}^\ast
Y_{J^{\prime\prime}-1, M^{\prime\prime}}\sin\theta d\theta d\phi\right)~,
\end{array}\end{equation}
where
\begin{equation}\label{6.15a}
a_{J,M}=\sqrt{\frac{(J+1)^2-M^2}{(2J+1)(2J+3)}}~.
\end{equation}
By the orthogonality of the spherical harmonics,
the above matrix elements vanish unless $J^{\prime\prime}=J^\prime\pm 1$.
The similar result is valid for
$R_x^{J^\prime M^\prime J^{\prime\prime} M^{\prime\prime}}$ and
$R_y^{J^\prime M^\prime J^{\prime\prime} M^{\prime\prime}}$. As a result the
selection rule of the emission (absorption) of the $q$-rotator model
is $\Delta J=\pm 1$. The emission (absorption) spectrum is
\begin{equation}\label{6.y1}\begin{array}{rcl}
\nu&=&\frac{\displaystyle E_{q-rot}(J+1)-E_{q-rot}(J)}{\displaystyle
hc},\\[2mm]
&=&B\left([J+1]_q[J+2]_q-[J+1]_q[J]_q\right),
\end{array}\end{equation}
where $B=\displaystyle\frac{h}{8\pi^2 Ic}$.
With an external electric field ${\bf F}$ an induced dipole moment
is formed.
Suppose that the external field is along $z$-axis,
and the induced dipole moment along the $z$-axis is
\begin{equation}\label{6.17a}
{\bf p}_z=\alpha_{zz}{\bf F}_z~,
\end{equation}
where $\alpha_{zz}$ is a component of the polarizability tensor
in the fixed frame. In terms of
the polarizability measured in the frame rotating with the molecules, it
is expressed by
\begin{equation}\label{6.18a}
\alpha_{zz}=\alpha_{x_mx_m}+
\left(\alpha_{z_mz_m}-\alpha_{x_mx_m}\right)\cos^2\theta~,
\end{equation}
where $\alpha_{x_{m}x_{m}}$ and $\alpha_{z_mz_m}$ are the components of the
polarizability tensor measured in the frame fixed on the rotating molecule.
The corresponding matrix elements are
\begin{equation}
\begin{array}{rcl}
\displaystyle\int \alpha_{zz}{\psi}^\ast_{J^\prime M^\prime}
{\psi}_{J^{\prime\prime}M^{\prime\prime}}d\tau
&=&\alpha_{x_mx_m}\displaystyle\int Y^\ast_{J^\prime M^\prime}
Y_{J^{\prime\prime}M^{\prime\prime}}d\tau\\[3mm]
& &+\left(\alpha_{z_mz_m}-\alpha_{x_mx_m}\right)\displaystyle
\int\cos^2\theta Y^\ast_{J^\prime M^\prime}
Y_{J^{\prime\prime} M^{\prime\prime}}d\tau~.
\end{array}
\end{equation}
According to the recursion relations of the spherical harmonics, the above
equation can be written as \\ \\
$\displaystyle\int \alpha_{zz}\tilde{\psi}^\ast_{J^\prime M^\prime}
\tilde{\psi}_{J^{\prime\prime}M^{\prime\prime}}d\tau$
\begin{equation}
\begin{array}{rcl}
&=&\left(\alpha_{x_mx_m}+\left(\alpha_{z_mz_m}-\alpha_{x_mx_m}
\right)\left(\left(a_{J^{\prime\prime},M^{\prime\prime}}\right)^2
+\left(a_{J^{\prime\prime}-1,M^{\prime\prime}}\right)^2\right)\right)
\displaystyle\int Y^\ast_{J^\prime M^\prime}
Y_{J^{\prime\prime}M^{\prime\prime}}d\tau\\[3mm]
& &+\left(\alpha_{z_mz_m}-\alpha_{x_mx_m}\right)
a_{J^{\prime\prime},M^{\prime\prime}}a_{J^{\prime\prime}+1,M^{\prime\prime}}
\displaystyle\int Y^\ast_{J^\prime M^\prime}
Y_{J^{\prime\prime}+2, M^{\prime\prime}}d\tau\\[3mm]
& &+\left(\alpha_{z_mz_m}-\alpha_{x_mx_m}\right)
a_{J^{\prime\prime}-1,M^{\prime\prime}}a_{J^{\prime\prime}-2,M^{\prime\prime}}
\displaystyle\int Y^\ast_{J^\prime M^\prime}
Y_{J^{\prime\prime}-2, M^{\prime\prime}}d\tau~.
\end{array}
\end{equation}
It is obvious that the first term in the above equation vanishes
unless $J^\prime=J^{\prime\prime}$,
i.e., it gives the lines without shifting; the second and third terms
vanish unless
$J^\prime=J^{\prime\prime}\pm 2$, i.e., it gives the shifted lines.
The same results can be obtained for the other two components.
Therefore the selection rule for the rotational Raman spectrum is
$\Delta J=\pm 2$. The rotational Raman spectrum can be expressed as
\begin{equation}\label{6.y2}
\begin{array}{rcl}
\nu&=&\nu_o+\frac{\displaystyle\left(E_{q-{\rm rot}}(J+2)-
E_{q-{\rm rot}}(J)\right)}{\displaystyle hc}~,\\[2mm]
&=&\nu_0+B\left([J+3]_{q}
[J+2]_{q}-[J+1]_{q}[J]_{q}\right)~,
\end{array}
\end{equation}
where $\nu_0$ is the wave number of the incident photon.
The rotational spectra of diatomic molecules involve the electronic
transitions and vibrational transitions, obeying the selection rule identical
to that for rigid rotator, i.e., $\Delta J=0,\pm 1$.
The rotational structures are
\begin{equation}\label{6.x0}
\begin{array}{clclll}
R&~{\rm branch:}&~
\nu&=&\nu_0+\frac{\displaystyle {E^\prime_{q-{\rm rot}}(J+1)-
E^{\prime\prime}_{q-{\rm rot}}(J)}}
{\displaystyle hc}~,\\[2mm]
&& &=&\nu_0+B^\prime[J+1]_{q^\prime} [J+2]_{q^\prime}
-B^{\prime\prime}[J]_{q^{\prime\prime}}
[J+1]_{q^{\prime\prime}}~;\\[4mm]
Q&~{\rm branch:}&~
\nu&=&\nu_0+\frac{\displaystyle {E^\prime_{q-{\rm rot}}(J)-
E^{\prime\prime}_{q-{\rm rot}}(J)}}
{\displaystyle hc}~,\\[2mm]
&& &=&\nu_0+B^\prime [J]_{q^\prime}[J+1]_{q^\prime}
-B^{\prime\prime}[J]_{q^{\prime\prime}}
[J+1]_{q^{\prime\prime}}~;\\[4mm]
P&~{\rm branch:}&~
\nu&=&\nu_0+\frac{\displaystyle {E^\prime_{q-{\rm rot}}(J-1)-
E^{\prime\prime}_{q-{\rm rot}}(J)}}
{\displaystyle hc}~,\\[2mm]
&& &=&\nu_0+B^\prime[J-1]_{q^\prime} [J]_{q^\prime}
-B^{\prime\prime}[J]_{q^{\prime\prime}}
[J+1]_{q^{\prime\prime}}~, \end{array} \end{equation} where $\nu_0$
is a quantity solely depending on the electronic transitions and
vibrational
structure, and $E^\prime_{q\rm -rot}$, $E^{\prime\prime}_{q\rm -rot}$ are
the eigenvalues of the $q$-rotator system in the upper and lower
electronic states respectively. Since internuclear distances are different
for different electronic states, the moments of inertia $I$'s of the
system are different for different electronic states. $B^\prime$
and $B^{\prime\prime}$ are different, because they are proportional to
$I^{-1}$.
\subsection{Vibrating-rotating structure}
The Hamiltonian describing the
vibrating-rotating structure of diatomic molecules \cite{ZC91c,ZC91d}
is
\begin{equation}\label{6.e388}
H_{q(J)-\rm vib}={1\over 2}\left(a_{q(J)}^\dagger a_{q(J)}+a_{q(J)}a_{q(J)}
^\dagger\right)hc\nu_{\rm vib}~.
\end{equation}
This Hamiltonian differs from that of the previous subsections, as
the quantization parameter $q$ is no longer a constant, but takes different
values at different rotational levels. The dependence of $q$ on the
rotational quantum number $J$ twists the effective rotational levels and
provides the necessary interaction of vibration and rotation.
It is obvious that this Hamiltonian commutes with the rotational one,
thus $H_{q(J)-{\rm vib}}$ and $H_{\rm rot}$ have common eigenstates.
The total Hamiltonian of the system reads
\begin{equation}
\begin{array}{rcl}
H_{q-\rm vib-rot}&=&H_{q(J)-{\rm vib}}+H_{\rm rot}\\[3mm]
&=&\displaystyle{1\over 2}\left(a_{q(J)}^\dagger
a_{q(J)}+a_{q(J)}a_{q(J)}^\dagger\right)hc
\nu_{\rm vib}+\displaystyle\frac{h^2}{8\pi^2I}C_I~,
\end{array}
\end{equation}
where $C_I$ is the Casimir operator
of the Lie group $SU(2)$. The dependence of $q$ on $J$ cannot be ignored
unless the interaction between vibration and rotation is negligible.
The quantum group $H_{q(J)}(4)$ is
\begin{equation}\label{6.27}
\begin{array}{l}
\displaystyle \left[a_{q(J)},a_{q(J)}^\dagger\right]=\left[N+1+b\gamma(J)
\right]_{q(J)}-\left[N+b\gamma(J)\right]_{q(J)}~,\\
\displaystyle \left[N, a_{q(J)}\right]=-a_{q(J)}~,~~~
\displaystyle \left[N, a_{q(J)}^\dagger\right]=a_{q(J)}^\dagger~,\\
\Delta(N')=N'\otimes 1+1\otimes N'-i\displaystyle\frac{\alpha}{\gamma(J)}
1\otimes 1~,\\
\Delta(a_{q(J)})=\left(a_{q(J)}\otimes q(J)^{N'/2}+iq(J)^{-N'/2}\otimes
a_{q(J)}\right)e^{-i\alpha/2}~,\\
\Delta(a_{q(J)}^\dagger)=\left(a_{q(J)}^\dagger\otimes q(J)^{N'/2}+
iq(J)^{-N'/2}\otimes a_{q(J)}^\dagger\right)e^{-i\alpha/2}~,\\
\epsilon(N')=\displaystyle i\frac{\alpha}{\gamma(J)}~,~~~~
\epsilon(a_{q(J)})=0=\epsilon(a_{q(J)}^\dagger)~,\\
S(N')=-N'+i\displaystyle\frac{2\alpha}{\gamma(J)}\cdot 1~,\\
S(a_{q(J)})=-q(J)^{-1/2}a_{q(J)}~,~~~~~S(a_{q(J)}^\dagger)=-q(J)^{1/2}
a_{q(J)}^\dagger~,
\end{array}
\end{equation}
where $N=a^\dagger a$, $N'=N+\displaystyle\frac{1}{2}+b\gamma(J)$.
It is interesting to note that all these structures are implicitly
$J$-dependent.
The Hamiltonian (\ref{6.e388}) can be cast into
\begin{equation}\label{6.28}
H_{{q(J)}-{\rm vib}}=\frac{1}{2}\left([N+b\gamma(J) ]_{q(J)}+[N+1+b\gamma(J)
]_{q(J)}\right)hc\nu_{\rm vib}~.
\end{equation}
The representation of $H_{q(J)}(4)$ in coordinate space is also
expressed by $H_v(x)$, the Hermite polynomial
\begin{equation}
\tilde{\psi}_{q(J)-\rm vib}(v,x)= N_v H_v(X)e^{-X^2/2}.
\end{equation}
The energy levels of the system are
\begin{equation}
E_{{q(J)}-{\rm vib}}(v,J)
={1\over 2}\left([v+1+b\gamma(J) ]_{q(J)}+[v+b\gamma(J) ]_{q(J)}\right)hc\nu_
{\rm vib}~.
\end{equation}
Thus the vibration-rotational spectrum of diatomic molecules has the form
\begin{equation}
E_{q-\rm vib-rot}(v,J)
={1\over 2}\left([v+1+b\gamma(J)]_{q(J)}+[v+b\gamma(J)]_{q(J)}\right)hc\nu_
{\rm vib}+\frac{h^2}{8\pi^2I}J(J+1)~.
\end{equation}
The vibration-rotational energy levels of a diatomic molecule at a certain
electronic state can be written as
\begin{equation}\begin{array}{rcl}
E&=&E_0+E_{q-\rm vib-rot}\\[2mm]
&=&E_0+\displaystyle \frac{\nu_{\rm vib}}{2}
\left(\left[v+b(J)\gamma(J)\right]_{q(J)}
+\left[v+b(J)\gamma(J)+1\right]_{q(J)}\right)hc\displaystyle
+\displaystyle\frac{h^2}{8\pi^2I} J(J+1)\\[3mm]
&=&E_0+\displaystyle {\nu_{\rm vib}}\frac{1}{2\sinh(\gamma(J)/2)}
\sinh\left(\gamma\left(J\right)\left(v+{1\over 2}+c(J)\right)\right)
+\displaystyle\frac{h^2}{8\pi^2I}J(J+1),
\end{array}
\end{equation}
where $E_0$ is the pure electron-transition energy and
$c(J)\equiv b(J)\gamma(J)$.
When $J=0$, there is no rotational excitation, and
\begin{equation}
\begin{array}{rcl}
E&=&E_0+E_{q(J)-{\rm vib}}(v)\\
&=&E_0+\displaystyle
\frac{1}{2\sinh(\gamma_0/2)}\sinh\left(\gamma_0\left(v+{1\over 2}+
c_0\right)\right)hc\nu_{\rm vib}~,
\end{array}
\end{equation}
which is just the vibrational spectrum.
When $v=0$, there is no vibrational excitation, and
\begin{equation}
\begin{array}{rcl}
E&=&E_0+E_{q(J)-{\rm vib}}(0)\\
&=&E_0+\displaystyle hc\nu_{\rm vib}\frac{1}{2\sinh(\gamma(J)/2)}
\sinh\left(\gamma
\left(J\right)\left({1\over 2}+c(J)\right)\right)\\[2mm]
& &+\displaystyle \frac{h^2}{8\pi^2I}J(J+1)~,
\end{array}
\end{equation}
which coincides in leading terms with the spectrum of the $q$-rotator
system. It should be noted that when $v=J=0$,
\begin{equation}
E=E_0+\frac{1}{2\sinh(\gamma(0)/2)}
\sinh\left(\gamma(0)\left(\frac{1}{2}+c(0)\right)\right)hc\nu_{\rm vib}
\end{equation}
which is $T_e$, the electronic term.
For simplicity assuming
\begin{equation}\label{6.e407}
\begin{array}{l}
\gamma(J)=\gamma_0+\gamma_1J(J+1)~,\\
c(J)=c_0+c_1J(J+1)~,
\end{array}\end{equation}
we have
\begin{equation}\label{6.e408}
\begin{array}{rcl}
E&=&E_0+\left(\displaystyle 2\sinh
\left({1\over2}\gamma_0+{1\over2}\gamma_1J(J+1)\right)
\right)^{-1}\times\\[3mm]
& &\times \sinh\left(\left(\displaystyle\gamma_0+\gamma_1J(J+1)\right)
\left(v+{\displaystyle 1\over\displaystyle
2}+c_0+c_1J(J+1)\right)\right)hc\nu_{\rm vib}\\[4mm]
& &+\displaystyle\frac{h^2}{8\pi^2I}J(J+1)~.
\end{array}
\end{equation}
This is the general form of the vibration-rotational energy levels of a
diatomic molecule. The second term represents the vibrational spectra in
interaction with the rotational, while the third describes the rigid rotation.
If we expand Eq.(\ref{6.e408}) into Taylor series, the parameters
$T_e$, $\omega_e$, $\omega_ex_e$, $\omega_ey_e$ and $\alpha_e$ introduced
in the conventional phenomenological treatment are reproduced as coefficients
of $\left(v+{1\over2}\right)^i\left(J\left(J+1\right)\right)^j$.
The total Hamiltonian for this system is
\begin{equation}
H=H_{\rm e}+H_{q(J)-{\rm vib}}+H_{\rm rot}~,
\end{equation}
which has the symmetry of $H_q(4)\otimes{\rm SU}(2)$.
The Hilbert space should be constructed from representations of the
symmetry $H_q(4)\otimes{\rm SU}(2)$, namely
\begin{equation}
\Psi_{q-\rm vib-rot}(v,J,x)=N_vH_v(X)e^{-X^2/2}Y_{JM}(\theta,\phi).
\end{equation}
The selection rule for infrared spectrum resulted from
$H_q(4)\otimes{\rm SU}(2)$ symmetry says that $v$ can change by arbitrary
integer
although $\Delta v=\pm 1$ gives the most intense transitions due to dipole
nature of the interaction, and $J$ can change only by $1$ due to the
observation of the total angular momentum. Of course $\Delta v=0$ is also
allowed, but this does not give rise to any rotation-vibrational spectrum but
the pure rotational one. If we now consider a particular transition from
$v^\prime$ to $v^{\prime\prime}$, the spectrum (in wavenumber) should be
\begin{equation}
\begin{array}{rcl}
\nu&=&\displaystyle\frac{\nu_{\rm vib}}{2}
\left[2\left(v^{\prime}+{1\over 2}+c_0+
c_1J^{\prime}(J^{\prime}+1)\right)
\right]_{ \left<\left(\gamma_0+\gamma_1J^{\prime}(J^{\prime}+1)
\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm
vib}}{2}\left[2\left(v^{\prime\prime}+{1\over 2}+c_0
+c_1J^{\prime\prime}(J^{\prime\prime}+1)\right)
\right]_{\left<\gamma_0+
\gamma_1J^{\prime\prime}(J^{\prime\prime}+1)/2\right>}\\[4mm]
& &+B_e\left(J^\prime(J^{\prime}+1)-J^{\prime\prime}(J^{\prime
\prime}+1)\right)~,
\end{array}\end{equation}
where the notation $[x]_{\left<\gamma\right>}= [x]_q$ is implied.
{}From the selection rule $\Delta J=1$ or $-1$, we have
\begin{equation}
\begin{array}{rcl}
\nu_R&=&\displaystyle \frac{\nu_{\rm vib}}{2}
\left[2\left(v^\prime+{1\over 2}+c_0+c_1(J+1)(J+2)\right)\right]
_{\left<\left(\gamma_0+\gamma_1(J+1)(J+2)\right)/2\right>}\\[4mm]
& &-\displaystyle \frac{\nu_{\rm vib}}{2}
\left[2\left(v^{\prime\prime}+{1\over 2}+c_0+c_1J(J+1)\right)\right]
_{\left<\left(\gamma_0+\gamma_1J(J+1)\right)/2\right>}\\[4mm]
& &+B_e\left((J+1)(J+2)-J(J+1)\right)~,
\end{array}
\end{equation}
and
\begin{equation}
\begin{array}{rcl}
\nu_P&=&\displaystyle \frac{\nu_{\rm vib}}{2}
\left[2\left(v^\prime+{1\over 2}+c_0+c_1J(J-1)
\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J-1)\right)/2\right>}\\[4mm]
& &-\displaystyle \frac{ \nu_{\rm vib}}{2}
\left[2\left(v^{\prime\prime}+{1\over 2}+c_0+c_1J(J+1)\right)\right]_
{\left<\left(\gamma_0+\gamma_1J(J+1)\right)/2\right>}\\[4mm]
& &+B_e\left(J(J-1)-J(J+1)\right)~,
\end{array}
\end{equation}
where $J^{\prime\prime}$ is replaced by $J$. Since $J$ can
take a whole series of values, these two formulae represent two series
of lines, which are called $R$, and $P$ branch respectively.
The selection rule for the Raman spectrum is $\Delta J=0,\pm 1$. Accordingly,
for a given Raman vibrational band,
there are three branches, for which the spectrum is readily obtained from
\begin{equation}
\begin{array}{rcl}
\Delta \nu&=&\displaystyle\frac{\nu_{\rm vib}}{2}\left[2\left(v^\prime
+{1\over 2}+c_0+c_1J^\prime (J^\prime +1)\right)\right]_
{\left<\left(\gamma_0+\gamma_1J^\prime (J^\prime +1)
\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm vib}}{2}\left[2\left
(v^{\prime\prime}+{1\over 2}+c_0+
c_1J^{\prime\prime}(J^{\prime\prime}+1)\right)\right]_{\left<
\left(\gamma_0+\gamma_1J^{\prime\prime}(J^{\prime\prime}+1)
\right)/2\right>}\\[4mm]
& &+B_e\left(J^\prime(J^{\prime}+1)-J^{\prime\prime}(J^{\prime
\prime}+1)\right),
\end{array}
\end{equation}
by substituting $J^\prime=J^{\prime\prime}+2$ for $S$ branch,
$J^\prime=J^{\prime\prime}-2$ for $O$ branch and $J^\prime=J^{\prime\prime}$
for $Q$ branch (and redenoting $J^{\prime\prime}=J$):
\begin{equation}
\begin{array}{rcl}
\left(\Delta \nu\right)_S&=&\displaystyle
\frac{\nu_{\rm vib}}{2}\left[2\left(v^{\prime}+{1\over 2}+
c_0+c_1(J+2)(J+3)\right)\right]_{\left<\left(\gamma_0+\gamma_1(J+2)(J+3)
\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm vib}}{2}\left[2\left(v^{\prime\prime}+{1\over
2}+
c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}\\[4mm]
& &+B_e(4J+6)~,
\end{array}
\end{equation}
where $J=0,1,\cdots$;
\begin{equation}
\begin{array}{rcl}
\left(\Delta \nu\right)_O&=&\displaystyle\frac{\nu_{\rm vib}}{2}
\left[2\left(v^{\prime}+{1\over 2}+
c_0+c_1(J-2)(J-1)\right)\right]_{\left<\left(
\gamma_0+\gamma_1(J-2)(J-1)\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm vib}}{2}\left[2\left(v^{\prime\prime}+
{1\over 2}+c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}\\[4mm]
& &+B_e(-4J+2)~,
\end{array}
\end{equation}
where $J=2,3,\cdots$;
\begin{equation}
\begin{array}{rcl}
\left(\Delta \nu\right)_Q&=&\displaystyle\frac{\nu_{\rm vib}}{2}
\left[2\left(v^{\prime}+{1\over 2}+
c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm vib}}{2}\left[2\left(v^{\prime\prime}+{1\over
2}+
c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}~,
\end{array}
\end{equation}
where $J=0,1,\cdots$.
Now we examine the vibrational-rotational structure of
elec\-tronic tran\-sitions, for which the wavenumber of the
transition is \begin{equation}
\nu=\left({E_0^\prime-E_0^{\prime\prime}}+{E_{q(J)-{\rm vib}}^\prime-
E_{q(J)-{\rm vib}}^{\prime\prime}}+{E_{rot}^\prime-
E_{\rm rot}^{\prime\prime}}\right)/hc~,
\end{equation}
where $E_0^\prime$, $E_{q(J)-{\rm vib}}^\prime$, $E_{\rm rot}^\prime$ and
$E_0^{\prime\prime}$, $E_{q(J)-{\rm vib}}^{\prime\prime}$, $E_{\rm
rot}^{\prime\prime}$
are the electronic energy and the vibration-rotational terms of the upper and
lower electronic state, respectively. The difference of the present spectra
from those of infrared and Raman lies in $E_{q(J)-{\rm vib}}^\prime$,
$E_{\rm rot}^\prime$ and $E_{q(J)-{\rm vib}}^{\prime\prime}$,
$E_{\rm rot}^{\prime\prime}$ belong to different electronic states
and have generally different magnitudes.
The selection rule tells us that the upper and lower states may have
different electronic angular
momenta $\Lambda$. If at least one of the two states has nonzero $\Lambda$,
the selection rule is $\Delta J=J^\prime-J^{\prime\prime}=0,\pm1$. However, if
$\Lambda=0$ in both electronic states, (i.e.,
$^1\Sigma\to~^1\Sigma$), the transition of $\Delta J=0$ is forbidden and
only the transitions of $\Delta J=\pm 1$ are allowed, as for most infrared
bands.
Expectedly, there are three or two series of lines (branches), for which the
wavenumbers are the following \\
$R$ branch:
\begin{equation}\begin{array}{rcl}
\nu&=&\displaystyle\frac{E_0^\prime-E_0^{\prime\prime}}{hc}+\displaystyle
\frac{\nu_{\rm vib}^\prime}{2}\left[2\left(v^{\prime}+{1\over 2}+
c_0+c_1(J+1)(J+2)\right)\right]_{\left<\left(
\gamma_0+\gamma_1(J+1)(J+2)\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm vib}^{\prime\prime}}
{2}\left[2\left(v^{\prime\prime}+{1\over 2}+
c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}\\[4mm]
& &+\left(B_e^\prime(J+2)(J+1)-B_e^{\prime\prime}J(J+1)\right)~;
\end{array}
\end{equation}
$Q$ branch:
\begin{equation}
\begin{array}{rcl}
\nu&=&\displaystyle\frac{E_0^\prime-E_0^{\prime\prime}}{hc}+\displaystyle
\frac{\nu_{\rm vib}^\prime}{2}\left[2\left(v^{\prime}+{1\over 2}+
c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm vib}^{\prime\prime}}{2}\left[2\left(
v^{\prime\prime}+{1\over 2}+
c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}\\[3mm]
& &+\left(B_e^\prime-B_e^{\prime\prime}\right)J(J+1)~;
\end{array}
\end{equation}
$P$ branch:
\begin{equation}
\begin{array}{rcl}
\nu&=&\displaystyle\frac{E_0^\prime-E_0^{\prime\prime}}{hc}+\displaystyle
\frac{\nu_{\rm vib}^\prime}{2}\left[2\left(v^{\prime}+{1\over 2}+
c_0+c_1(J-1)J\right)\right]_{\left<\left(\gamma_0+\gamma_1(J-1)J
\right)/2\right>}\\[4mm]
& &-\displaystyle\frac{\nu_{\rm vib}^{\prime\prime}}
{2}\left[2\left(v^{\prime\prime}+{1\over 2}+
c_0+c_1J(J+1)\right)\right]_{\left<\left(\gamma_0+\gamma_1J(J+1)
\right)/2\right>}\\[3mm]
& &+\left(B_e^\prime(J-1)J-B_e^{\prime\prime}J(J+1)\right)~.
\end{array}
\end{equation}
This completes the description for vibrating-rotating
diatomic molecules in the quantum group theoretic approach.
\bigskip\bigskip\bigskip
\centerline{\bf Acknowledgments}
I would like to thank Prof. H. Y. Guo for his encouragement to
prepare this Review. The author is indebted to Prof. Yang-Zhong Zhang for
his reading the manuscript and helps in rhetoric.
|
\section{Introduction}
\label{sec:intro}
Renormalisation of lattice operators is a necessary step for obtaining
physical
results from numerical simulations. In this paper, we apply
the general method introduced in \cite{NP} to the four-fermion
operator\footnote{We use the Euclidean metric throughout this
paper.} \begin{equation}
O^{\Delta S=2}=(\bar s \gamma_{\mu}^L d )(\bar s \gamma_{\mu}^L d)
\, , \label{eq:O_DS=2} \end{equation}
which appears in the weak effective Hamiltonian relevant
for $K^0$--$\bar K^0$ mixing
\begin{equation}
{\cal H}^{\Delta S=2}_{\mbox{\small eff}}=C(M_W/\mu) O^{\Delta S=2}(\mu) \, ,
\end{equation}
where $\gamma_{\mu}^L=\frac{1}{2}\gamma_{\mu} (1-\gamma_5)$,
$O^{\Delta S=2}(\mu)$ is the renormalised operator,
$C(M_W/\mu)$ is
the corresponding Wilson coefficient and $\mu$ the renormalisation scale.
The $K^0$--$\bar K^0$ matrix element of $O^{\Delta S=2}(\mu) $
defines the so-called kaon $B$-parameter
\begin{equation}
\<\bar K^0| O^{\Delta S=2}(\mu) | K^0\>=\frac{8}{3}f_K^2m_K^2B_K(\mu)\, .
\label{eq:B_K}
\end{equation}
The uncertainty in the value of this matrix element restricts the
precision with which the CKM matrix elements $\rho$ and $\eta$ (in
the Wolfenstein parametrisation) can be determined from experimental
measurements. It is therefore of considerable importance to determine
this matrix element using lattice simulations.
In the continuum, chiral symmetry implies that
the kaon matrix element of $O^{\Delta S=2}$ vanishes in the chiral
limit \cite{Cabibbo,Gellmann}
\begin{equation}
\<\bar K^0(q) | O^{\Delta S=2}(\mu) | K^0(p)\>=
\gamma (p\cdot q) + O\left( (p\cdot q)^2 \right) \, .
\label{eq:B_K_chiral} \end{equation}
On the lattice however, in simulations based on Wilson's formulation of
the fermion action
(such as the standard Wilson action or the SW-Clover action \cite{sw}),
the presence of chiral symmetry breaking terms leads to the mixing of
$O^{\Delta S=2}$ with operators of different chirality
\cite{marti84}--\cite{improved}, and the matrix element of $O^{\Delta
S=2}$ is different from zero at $p \cdot q=0$
\cite{bern}--\cite{Crisafulli}. For this reason, it is possible to
define a renormalised operator with definite chiral properties only in
the continuum limit, i.e. when $a \to 0$. At finite $a$, one can
improve the chiral behaviour of the matrix element of $O^{\Delta S=2}$,
by sub\-tracting a suitable set of dimension six operators. The mixing
coefficients have so far been computed only in one-loop perturbation
theory \cite{marti84},\cite{berw}--\cite{improved}. In this way, the
systematic error in the value of the matrix element determined on the
lattice is of $O(\alpha_s^2)$. In addition, as a consequence of the
finiteness of the lattice spacing, there are errors of $O(a)$.
Following Symanzik's proposal, one can reduce these discretization
errors from $O(a)$ to $O(\alpha_s a)$ by using the tree-level
``improved'' SW-Clover lattice quark action \cite{sw,clover}. Using
this action, the improvement has been shown to be effective
for two-fermion operators, at values of $\beta$ currently used in
numerical simulations \cite{msv}--\cite{wiukqcd}. It remains true
however, that ignorance of higher-order perturbative corrections to the
mixing coefficients can distort the chiral behaviour of the operator
and hence induce a large systematic error in the determination of
$B_K$. The use of a non-perturbative approach to the determination of
the renormalisation constants, should reduce this systematic effect.
In the following, we will define a renormalised operator $O^{\Delta
S=2}(\mu)$, obtained by applying the non-perturbative method of ref.
\cite{NP} to the computation of the mixing coefficients and of the
overall renormalisation constant.
In order to reduce the discretisation errors, including those induced
by the mixing with higher dimensional operators, it is necessary to
use an improved fermion action and operators. In the computations
described below we have used the improved SW-Clover action and the
``improved-improved'' operators introduced in ref.
\cite{improved,tass2}. We monitor the effects of the
non-perturbative determination of the mixing coefficients by comparing
the chiral behaviour of the matrix element $\<\bar K^0(q)|O^{\Delta
S=2}|K^0(p)\>_{\mbox{ latt}}$ computed by using the operator renormalised
with standard or boosted \cite{Lepage} perturbation theory to the
matrix element of the operator renormalised non-perturbatively. In
particular, by parametrizing the matrix element near the chiral limit
in the standard way \cite{capri}--\cite{gupta},
\begin{equation}
\<\bar K^0|O^{\Delta S=2}|K^0\>_{\mbox{ latt}}=\alpha+\beta m_K^2+\gamma (p\cdot q)
+\delta m_K^4+\epsilon m_K^2(p\cdot q) +\zeta (p\cdot q)^2+\ldots ,
\label{eq:B_K_latt}\end{equation}
we investigate the differences in the values of
$\alpha$ and $\beta$, obtained by fitting the dependence of the
matrix element on the kaon masses and momenta.
Since in the continuum
$\alpha$ and $\beta$ are absent, cf. eq.\ (\ref{eq:B_K_chiral}), we consider a
reduction of their values as a measure of the improvement in the chiral
behaviour and in the accuracy of the determination of the matrix element.
Using the data of the APE collaboration \cite{Donini94,Crisafulli},
we show that the chiral behaviour is indeed improved by
using the non-perturbative results.
The paper is organized as follows. In section \ref{sec:strategy}, we
briefly summarize the strategy followed for computing the mixing
coefficients and the overall renormalisation constant of the relevant
four-fermion operator; in section \ref{sec:mixing}, we illustrate the
projection method used to determine the mixing coefficients.
Although the method is applied specifically to the renormalisation
of the operator $O^{\Delta S=2}$, it can readily be generalised
to other sets of operators which mix under renormalisation.
In section~\ref{sec:PT} we give some information about the perturbative
evaluation of the renormalisation constants on the lattice;
in section~\ref{sec:numerical} we present the details of the numerical
simulation and discuss our results and, finally, we present our conclusions
in section~\ref{sec:conclusion}.
\section{The non-perturbative method for four-fer\-mion operators}
\label{sec:strategy}
The renormalisation method proposed in \cite{NP} completely avoids the
use of lattice perturbation theory and allows for a non-perturbative
determination of the renormalisation constants of any composite
operator in a renormalisation scheme which is independent of the method
used to regulate the ultra-violet divergences. In particular, the
renormalised operators are independent of the fact that we start from
bare operators in lattice QCD. To stress this point further we will
refer to the renormalisation scheme defined below for $O^{\Delta S=2}$
as the RI (Regularization Independent) scheme \cite{Ciuchini2}\footnote{
Although of course such
a name could be applied equally well to many other schemes.}.
Non-perturbative renormalisation conditions are imposed directly on
quark Green functions with off-shell external states in a fixed gauge,
for example the Landau gauge. The method is expected to work in all
cases where it is possible to fix the virtuality of the external states
$p^2=\mu^2$ so as to satisfy the condition $\Lambda_{{\rm QCD}}\ll \mu \ll
1/a$. The condition $\mu \gg \Lambda_{{\rm QCD}}$ is necessary because one
has to match perturbatively the effective Hamiltonian, expressed in
terms of operators renormalised at the scale $\mu$, to the full theory.
This condition is common to all approaches currently used. The
condition $\mu \ll 1/a$ is a requirement common to all lattice methods
and is due to the presence of $O(a)$ ($O(\alpha_s a)$) effects in the
operator matrix elements. The existence of the ``window"
$\Lambda_{{\rm QCD}}\ll \mu \ll 1/a$ depends on the value of the bare
lattice coupling $\beta$ at which the numerical calculations are
performed. We refer the reader to ref. \cite{NP} for a more detailed
discussion on this point.
In the following, in order to use a more transparent notation in the formulae,
we will consider the operator
\begin{equation}
O_+= \frac{1}{2}
[(\bar\psi_1\gamma^L_{\mu}\psi_2)(\bar\psi_3\gamma^L_{\mu}\psi_4)
+(2\leftrightarrow 4)],
\label{eq:O_+} \end{equation}
with four distinct quark flavours ($f=1,2,3,4$) instead of the operator
$O^{\Delta S=2}$. $O_+$ and $O^{\Delta S=2}$ have the same
renormalisation properties.
The discretization of the quark action {\em \`a la} Wilson, induces a
mixing of the operator (\ref{eq:O_+}) with operators of a different
chirality which, in the language of refs.
\cite{Ciuchini2}--\cite{Ciuchini}, correspond to the so-called
``effervescent'' (``evanescent'') operators. The mixing, being a
consequence of the regularization procedure, is not limited to the
lattice case, but is present also in continuum regularizations. The
effervescent operators must be subtracted from the bare one by a
suitable renormalisation procedure \footnote{ Using dimensional
regularization, the one-loop mixing with the ``effervescent''
operators is cancelled by the minimal subtraction of the pole in
$1/\epsilon$.}.
In the lattice case, CPS symmetry
fixes the basis of operators that may
appear in perturbation theory \cite{BERNARD2}:
\begin{eqnarray}
O^{SP}_+&=&-\frac{1}{16N_c}
[(\bar\psi_1\psi_2)(\bar\psi_3\psi_4)
-(\bar\psi_1\gamma_5\psi_2)(\bar\psi_3\gamma_5\psi_4)
+(2\leftrightarrow 4) ], \label{eq:O_+^SP} \\
O^{VA}_+&=&-\frac{(N_c^2+N_c-1)}{32N_c}
[(\bar\psi_1\gamma_{\mu}\psi_2)(\bar\psi_3\gamma_{\mu}\psi_4)
-(\bar\psi_1\gamma_{\mu}\gamma_5\psi_2)(\bar\psi_3\gamma_{\mu}\gamma_5\psi_4)
\nonumber \\
& &\qquad +(2\leftrightarrow 4) ], \label{eq:O_+^VA} \\
O^{SPT}_+&=&\frac{(N_c-1)}{16N_c}
[(\bar\psi_1\psi_2)(\bar\psi_3\psi_4)
+(\bar\psi_1\gamma_5\psi_2)(\bar\psi_3\gamma_5\psi_4) \nonumber \\
& & \qquad
+(\bar\psi_1\sigma_{\mu\nu}\psi_2)(\bar\psi_3\sigma_{\mu\nu}\psi_4)
+(2\leftrightarrow 4) ], \label{eq:O_+^SPT}
\end{eqnarray}
where $N_c=3$ denotes the number of colours.
We renormalise the operator $O_+$ by introducing the subtracted
operator $O^s_+$,
\begin{equation}
O_+(\mu)= Z_+ O^s_+=Z_+
(O_+ +Z_1 O^{SP}_+ +Z_2 O^{VA}_+ +Z_3 O^{SPT}_+)\, ,
\label{eq:O_+(mu)}\end{equation}
where $Z_{+,1,2,3}=Z_{+,1,2,3}(\mu a , g_0^2(a))$ and the bare lattice
coupling is given by the relation $\beta= 6/ g_0^2(a)$. The mixing
constants $Z_i,\ i=1,\ldots,3$, are determined by means of projection
operators that will be defined in sec\-tion~\ref{sec:mixing}. Their
values are fixed by the requirement that, up to terms of $O(\alpha_s
a)$, $O^s_+$ renormalises multiplicatively. $O^s_+$ is logarithmically
divergent as $a\to 0$, and this divergence is removed by imposing a
renormalisation condition on $O^s_+$ which defines the overall
renormalisation constant $Z_+(\mu a,g^2_0(a))$,
\begin{equation}
Z_+(\mu a,g^2_0(a))Z_{\psi}^{-2}(\mu a,g^2_0(a))
\Gamma^s_+(pa)|_{p^2=\mu^2}=1,
\label{eq:Z_+}\end{equation}
where $\Gamma^s_+(pa)$ is obtained by projecting a suitable amputated
Green function of the operator $O^s_+$ on the Dirac structure
$\gamma^L_{\mu}\otimes \gamma^L_{\mu}$ (see eq.\ (\ref{eq:P_0}) in
section \ref{sec:mixing} and refs.\ \cite{NP,Ciuchini2}). $Z_{\psi}$ is
the quark field renormalisation constant to be defined below (eq.\
(\ref{eq:Z_psi}) of section \ref{sec:mixing} and ref. \cite{NP}). In
eq.\ (\ref{eq:Z_+}), $p^2=\mu^2$ denotes the momentum of the external
quark states. We have chosen equal momenta for all four external quark
legs, because this is the simplest choice which regulates the infrared
divergences \cite{Ciuchini2}.
The renormalised operator in eq.\ (\ref{eq:O_+(mu)}), calculated in the
RI scheme, depends both on the gauge and on the external states. The
Wilson coefficient must also be calculated in the same gauge and with
the same external states in order to obtain the physical operators
which are independent of both\footnote{This is true up to higher order
continuum perturbative corrections and lattice systematic errors.}.
The next-to-leading order calculation of the Wilson coefficient
relevant for the operator (\ref{eq:O_+}), in the Landau gauge and with
equal external momenta, can be found in ref. \cite{Ciuchini2}.
\section{Determination of the mixing constants}
\label{sec:mixing}
In this section, we define the four-point amputated Green functions and
introduce the projectors that have been used to determine the mixing
constants.
Since the non-perturbative renormalisation conditions are imposed on
quark states, the Green functions of a four-fermion operator will
depend on four coordinates. Denoting by $x_1,x_3$ and $x_2,x_4$ the
coordinates of the outgoing and incoming quarks, the Green functions
corresponding to the insertion of the operators
(\ref{eq:O_+})--(\ref{eq:O_+^SPT}) can be written as linear
combinations of Green functions of the form
\begin{equation}
G_{\Gamma^a}(x_1,x_2,x_3,x_4)=\<\psi_1(x_1)\bar\psi_2(x_2) O_{\Gamma^a}(0)
\psi_3(x_3)\bar\psi_4(x_4)\>\, , \label{eq:G_Gamma(x)}
\end{equation}
where $\<\cdots\>$ denotes the vacuum expectation value, i.e. the
average over the gauge-field configurations.
The generic four-fermion operator $O_{\Gamma^a}$ is given by
\begin{equation}
O_{\Gamma^a}(0)=C_{\Gamma^a}\left[
\bar\psi_1(0)\Gamma^a\psi_2(0) \bar\psi_3(0)\Gamma^a\psi_4(0)
+\bar\psi_1(0)\Gamma^a\psi_4(0) \bar\psi_3(0)\Gamma^a\psi_2(0)
\right]\, ,
\label{eq:O_Gamma}\end{equation}
where $\Gamma^a$ denotes a Dirac matrix, and $C_{\Gamma^a}$ is a
constant associated with $\Gamma^a$. The index $a$ can be either
single-valued (if $\Gamma^a = \mbox{1$\!\!$I}$ or $\gamma_5$) or be summed
over a range of values (if $\Gamma^a = \gamma_{\mu}$,
$\gamma_{\mu}\gamma_5$ or $\sigma_{\mu\nu}$ a sum over repeated Lorentz
indices is implied).
The Fourier transform of the non-amputated Green function
(\ref{eq:G_Gamma(x)}), at equal external momenta $p$, has the form
\begin{equation}
G_{\Gamma^a}(p)^{ABCD}_{\alpha\beta\gamma\delta}
=C_{\Gamma^a}\left[
\<\Gamma^a(p)^{AB}_{\alpha\beta}\otimes \Gamma^a(p)^{CD}_{\gamma\delta}\>
-\<\Gamma^a(p)^{AD}_{\alpha\delta}\otimes\Gamma^a(p)^{CB}_{\gamma\beta}\>
\right]\, ,
\label{eq:G_Gamma(p)}\end{equation}
where
\begin{equation}
\Gamma^a(p)^{XY}_{\chi\psi}=S(p|0)^{XR}_{\chi\rho}\Gamma^a_{\rho\sigma}
(\gamma_5S(p|0)^{\dag}\gamma_5)^{RY}_{\sigma\psi}\, .
\label{eq:gammaa} \end{equation}
In eqs. (\ref{eq:G_Gamma(p)}) and (\ref{eq:gammaa}) the upper-case
Roman superscripts denote colour labels and the lower case Greek
subscripts denote spinor labels. $S(p|0)$ is defined by
\begin{equation}
S(p|0)=\int d^4x S(x|0) e^{-ip\cdot x},
\end{equation}
where $S(x|0)$ is the quark propagator computed on a single gauge-field
configuration (cf. section 4 of \cite{NP}), and is therefore not
translationally invariant. It satisfies the relation
\begin{equation}
S(x|0)=\gamma_5 S^{\dag}(0|x)\gamma_5.
\end{equation}
The amputated Green function can be obtained from eq.\
(\ref{eq:G_Gamma(p)})
\begin{equation}
\Lambda_{\Gamma^a}(p)^{RSR'S'}_{\rho\sigma\rho'\sigma'}
=S^{-1}(p)^{RA}_{\rho\alpha}S^{-1}(p)^{R'C}_{\rho'\gamma}
G_{\Gamma^a}(p)^{ABCD}_{\alpha\beta\gamma\delta}
S^{-1}(p)^{BS}_{\beta\sigma}S^{-1}(p)^{DS'}_{\delta\sigma'}\, ,
\label{eq:Lambda_Gamma(p)}\end{equation}
where $S(p)$ is the Fourier transform of the translationally-invariant
quark propagator, i.e. the Fourier transform of $S(x|0)$, averaged over
the gauge-field configurations.
As mentioned above, the renormalisation procedure necessary to
determine the mixing constants consists in defining suitable
projectors on the amputated Green functions of the operators
(\ref{eq:O_+})--(\ref{eq:O_+^SPT}). To this end let us introduce a more
convenient notation. Let us denote by $O_i,\ i=0,\ldots,3$,
respectively, the operators $O_+,O_+^{SP},O_+^{VA},O_+^{SPT}$. Then,
the projectors $\Pr_i,\ i=0,\ldots,3$, are defined by the condition
\begin{equation}
\mbox{Tr } \Pr_i \Lambda^{(0)}_j=\delta_{ij},\qquad i,j=0,\ldots,3, \label{eq:P_i}
\end{equation}
where $\Lambda^{(0)}_i,\ i=0,\ldots,3$, are the amputated Green
functions, at tree level, of the operators $O_i$, and the trace is
understood over colour and spin (as defined below).
The renormalisation scheme depends on the precise definition of the
projection operators and we now define our procedure in detail.
For each Dirac Matrix $\Gamma^b$, we define the projector
$\Pr_{\Gamma^b}$ by~\footnote{It is only the traces
(\ref{eq:proj_def_1}) which are required for the determination of the
subtraction constants.}
\begin{eqnarray}
\mbox{Tr }\Pr_{\Gamma^b} \Lambda_{\Gamma^a}(p)
=(\Gamma^b_{\sigma\rho}\otimes \Gamma^b_{\sigma'\rho'})
\Lambda_{\Gamma^a}(p)^{RRR'R'}_{\rho\sigma\rho'\sigma'}\, ,
\label{eq:proj_def_1}
\end{eqnarray}
where the index $b$ is either fixed or corresponds to a sum over
repeated indices. In the free theory, the amputated Green function
reduces to
\begin{equation}
\Lambda^{(0)}_{\Gamma^a}(p)^{RSR'S'}_{\rho\sigma\rho'\sigma'}
=C_{\Gamma^a}[
\delta^{RS}\delta^{R'S'}(\Gamma^a_{\rho\sigma}\otimes\Gamma^a_{\rho'\sigma'})
-\delta^{RS'}\delta^{R'S}(\Gamma^a_{\rho\sigma'}\otimes\Gamma^a_{\rho'\sigma})].
\label{eq:Lambda_+_0}\end{equation}
and the result of the projection defined in (\ref{eq:proj_def_1}) is:
\begin{eqnarray}
\mbox{Tr } \Pr_{\Gamma^b} \Lambda^{(0)}_{\Gamma^a}(p)
=C_{\Gamma^a}[
N_c^2 (\mbox{Tr } \Gamma^a \Gamma^b)(\mbox{Tr } \Gamma^a \Gamma^b)
- N_c (\mbox{Tr } \Gamma^a\Gamma^b\Gamma^a\Gamma^b)].
\label{eq:proj_free_1}\end{eqnarray}
The projectors corresponding to $O_+$ and to the operators $O_i$ defined in
eqs.(\ref{eq:O_+^SP})--(\ref{eq:O_+^SPT}) are as follows:
\begin{eqnarray}
\Pr_0&=&\frac{1}{8N_c(N_c+1)}\Pr_{\gamma^R_{\mu}} , \label{eq:P_0}\\
\Pr_1&=&\frac{N_c}{2(1-N_c^2)}(\Pr_{\mbox{\small 1$\!\!$I}}-\Pr_{\gamma_5}) \nonumber \\
&+&\frac{1}{4(1-N_c^2)}(\Pr_{\gamma_{\mu}}-\Pr_{\gamma_{\mu}\gamma_5}),
\label{eq:P_1}\\
\Pr_2&=&\frac{1}{2(1-N_c^2)(N_c^2+N_c-1)}(\Pr_{\mbox{\small 1$\!\!$I}}-\Pr_{\gamma_5}) \nonumber \\
&+&\frac{N_c}{4(1-N_c^2)(N_c^2+N_c-1)}
(\Pr_{\gamma_{\mu}}-\Pr_{\gamma_{\mu}\gamma_5}) \label{eq:P_2}\\
\Pr_3&=&\frac{1}{8(N_c^2-1)}(\Pr_{\mbox{\small 1$\!\!$I}}+\Pr_{\gamma_5}+\Pr_{\sigma_{\mu\nu}}),
\label{eq:P_3}\end{eqnarray}
where $\gamma^R_{\mu}=\frac{1}{2}\gamma_{\mu}(1-\gamma_5)$. Note that
the projectors $\Pr_0$ and $\Pr_3$, eqs. (\ref{eq:P_0}) and
(\ref{eq:P_3}), have the same Dirac structure as the operators $O_0$
and $O_3$, eqs. (\ref{eq:O_+}) and (\ref{eq:O_+^SPT}). This is due to
the Fierz rearrangement properties of these operators.
It is possible to determine the mixing coefficients
$Z_i$ by using the projectors (\ref{eq:P_0})--(\ref{eq:P_3}) defined in
the free field case. Let us introduce the matrix $D$ defined by
\begin{equation} \Lambda_i=\sum_{j=0}^3 D_{ij} \Lambda^{(0)}_j
\, , \end{equation}
where the elements $D_{ik}$ are determined non-perturbatively by the
projections
\begin{equation}
D_{ik}=\mbox{Tr } \Pr_k \Lambda_i,\qquad i,k=0,\ldots,3.
\end{equation}
The mixing constants $Z_i$ are then fixed by the condition that the
subtracted operator $O_+^s$ is proportional to the bare free
operator
\begin{equation}
\mbox{Tr } \Pr_k \Lambda^s_+=
\left(D_{0k}+\sum_{i=1}^3Z_i D_{ik} \right)=0,\qquad k=1,2,3\, .
\label{eq:mixing_condition}
\end{equation}
Equation (\ref{eq:mixing_condition}) yields three conditions
corresponding to a linear non-homo\-geneous system in the three unknowns
$Z_i$. Defining the reduced $3\times 3$ matrix $\tilde D$ as
\begin{equation}
\tilde D_{ik}=D_{ki},\qquad i,k=1,2,3 \, ,
\end{equation}
the solutions of this linear system are given by
\begin{equation}
Z_i=-\sum_{k=1}^3(\tilde D)^{-1}_{ik}D_{0k},\qquad i=1,2,3 \, .\label{eq:Z_i}
\end{equation}
The overall renormalisation constant $Z_+$ is then determined by the condition
(\ref{eq:Z_+}), using
\begin{equation}
\Gamma^s_+(pa)= \mbox{Tr } \Pr_0 \Lambda^s_+
=\left(D_{00}+\sum_{i=1}^3 Z_i D_{i0}\right).
\end{equation}
In eq.\ (\ref{eq:Z_+}),
the renormalisation constant $Z_{\psi}$ is defined by the relation
\begin{equation}
Z_{\psi}(\mu a, g_0^2(a))=
\left.\frac{1}{48}\mbox{Tr } \left( \Lambda_{V^L_{\mu}}\gamma_{\mu}\right)
\right|_{p^2=\mu^2}\times Z_{V^L}, \label{eq:Z_psi}
\end{equation}
where $V^L=\bar\psi \gamma_{\mu}\psi$ is the local vector current, and
$Z_{V^L}$ its renormalisation constant which can be determined
with high accuracy, by using the vector current Ward identities
\cite{wi,wiukqcd,mm}. $\Lambda_{V^L_{\mu}}$ is defined as
\begin{equation}
\Lambda_{V^L_{\mu}}(p)=S(p)^{-1}G_{V^L_{\mu}}(p)S(p)^{-1}\, ,
\end{equation}
where $G_{V^L_{\mu}}(p)$ is the non-amputated two-point Green function
of the local vector current, $G_{V^L_{\mu}}(p)=\langle
\Gamma^{V^L_{\mu}}(p)\rangle$, cf. eq.\ (\ref{eq:gammaa}). There are
various equivalent ways to define $Z_{\psi}$, but (\ref{eq:Z_psi}) is
the most natural from a non-perturbative point of view. For a more
thorough discussion on the determination of $Z_{\psi}$, we refer the
reader to sections 2 and 4 of ref. \cite{NP}.
\section{Lattice perturbation theory}
\label{sec:PT}
We have also calculated the renormalisation constants $Z_+, Z_1, Z_2$
and $Z_3$ in one-loop perturbation theory, in order to be able to
compare the results with those obtained non-perturbatively. Since the
non-perturbative renormalisation condition depends on the gauge and on
the external states, the perturbative calculation must be done in the
Landau gauge and at equal external momenta. This calculation is an
extension of those of refs. \cite{4f,improved}.
Starting from a bare lattice operator $O(a)$, the one-loop vertex
function $\Gamma_O^{\lambda}(pa)$ is obtained by tracing the amputated
Green function
(but with wave function effects included) with a suitable projector.
The generic expression of
$\Gamma_O^{\lambda}(pa)$, calculated between states of
momentum $p$ and in a fixed gauge $\lambda$, is
\begin{equation}
\Gamma_O^{\lambda}(pa)=
\left[1+\frac{\alpha_s}{4\pi}\left(\gamma_O\log(1/p^2 a^2)+
r_O^{\mbox{ latt}}(\lambda,p)
\right)\right] \, .
\end{equation}
$\gamma_O$ is the anomalous dimension, which at one-loop order is
independent of the gauge, the external states and the
regularization\footnote{ $\lambda$ denotes a generic covariant gauge:
$\lambda=0$ corresponds to the Landau gauge,
$\lambda=1$ correspond to the Feynman gauge.}.
The finite coefficient $r_O^{\mbox{ latt}}(\lambda,p)$, on the other hand, does
depend on the gauge, the regularization and the external states.
The momentum label $p$ appearing as argument of
$r_O^{\mbox{ latt}}(\lambda,p)$ indicates that the result is a dimensionless
function of the external states.
In the continuum, in any renormalisation scheme based on dimensional
regularization (DR=NDR, HV or DRED),
the vertex function between states of momentum $p$
and in a generic gauge, is
\begin{equation}
\Gamma_O^{\lambda}\left(p/\mu\right)=
\left[1+\frac{\alpha_s}{4\pi}\left(\gamma_O\log(\mu^2/p^2)+r_O^{{\rm DR}}(\lambda,p)
\right)\right]\, ,
\end{equation}
where $\mu$ is the DR renormalisation scale. Thus, the one-loop
relation between the operators in the continuum and on the lattice is
\begin{equation}
O(\mu)=\left[1+\frac{\alpha_s}{4\pi}
\left(\gamma_O\log(\mu^2a^2)+\Delta^{{\rm DR}-\mbox{ latt}}\right)\right] O(a),
\end{equation}
where
\begin{equation}
\Delta_O^{{\rm DR}-\mbox{ latt}}=r_O^{{\rm DR}}(\lambda,p)-r_O^{\mbox{ latt}}(\lambda,p)
\label{eq:Delta_DR_Latt}
\end{equation}
is independent of both $\lambda$ and $p$. From
$\Delta_O^{{\rm DR}-\mbox{ latt}}$, we can calculate the lattice constant
$r_O^{\mbox{ latt}}(\lambda,p)$, in any gauge and at any external momenta,
from the corresponding constant in the continuum,
$r_O^{{\rm DR}}(\lambda,p)$. In order to compare the perturbative result
with the non-perturbative determination, we need
$r_O^{\mbox{ latt}}(\lambda=0,p)$ in the Landau gauge and with
non-zero but equal external momenta. From (\ref{eq:Delta_DR_Latt}), we
immediately obtain
\begin{equation}
r_O^{\mbox{ latt}}(\lambda=0,p)=r_O^{{\rm DR}}(\lambda=0,p)
-\Delta_O^{{\rm DR}-\mbox{ latt}}.
\end{equation}
Since $r_O^{\mbox{ latt}}$ must be independent of the continuum regularization used
in the intermediate steps,
a check of the correctness of the calculation is given by
\begin{equation}
r_O^{{\rm DRED}}(\lambda=0,p)-\Delta_O^{{\rm DRED}-\mbox{ latt}}
=r_O^{{\rm NDR}}(\lambda=0,p)-\Delta_O^{{\rm NDR}-\mbox{ latt}},
\end{equation}
which is equivalent to
\begin{equation}
r_O^{{\rm DRED}}(\lambda=0,p)-r_O^{{\rm NDR}}(\lambda=0,p) =
r_O^{{\rm DRED}}(\lambda=1,p^\prime )-r_O^{{\rm NDR}}(\lambda=1,p^\prime ).
\end{equation}
The one-loop contribution to the renormalisation constant $Z_+$ and to
the mixing coefficients $Z_i$'s have been calculated in
\cite{4f,improved}, by comparing the lattice and the DRED scheme. In
the notation of these authors
\begin{eqnarray}
Z_+&=&1 + \frac{\alpha_s}{4 \pi} F_+,\qquad F_+=\Delta_{O_+}^{{\rm DRED}-\mbox{ latt}}
=-10.9\, ,\nonumber \\
Z_1&=&Z_2=Z_3=\frac{\alpha_s}{4 \pi} F^*, \qquad F^{\ast}=19.4\, .
\label{contlatt2}
\end{eqnarray}
In the DRED scheme, for a generic gauge $\lambda$ and by taking the
momenta of the external legs to be equal, we find
\begin{equation}
r_{O_+}^{{\rm DRED}}(\lambda,p)=\lambda \Bigl( - 7/3 + 8/3\log(2) \Bigr)
- 5/3 + 8\log(2) \, , \label{eq:r_+_DRED}
\end{equation}
whilst the mixing with the ``effervescent" operators is
cancelled by the minimal subtraction procedure.
Thus, the perturbative
expressions of the lattice renormalisation constants
in the RI scheme are
given by\footnote{ The scale $\mu$ in this formula denotes the
renormalization scale at which $\Gamma_O^{\lambda}(pa)$ is renormalized,
i.e. $Z_+ \Gamma_O^{\lambda}(pa)\vert_{p^2=\mu^2}=1$.}
\begin{eqnarray}
Z_+^{{\rm RI}}&=&1-\frac{\alpha_s}{4\pi}\left(-\gamma_{O_+}\log(\mu^2a^2)
+r_{O_+}^{{\rm DRED}}(\lambda=0,p)-F_+\right),
\nonumber \\ Z_i^{{\rm RI}}&=&\frac{\alpha_s}{4\pi}F^{\ast},
\end{eqnarray}
with $\gamma_{O_+}= - 2$ and $r_{O_+}^{{\rm DRED}}(\lambda,p)$ given in
eq.\ (\ref{eq:r_+_DRED}).
We have not evaluated the renormalisation constants using Discrete
Perturbation Theory (DPT), i.e. by summing only over the discrete
values of momenta allowed on our finite lattice (this was done for the
two-quark operators in ref. \cite{NP}). We have only evaluated the
constants using standard lattice perturbation theory, in which finite
lattice size effects are neglected.
In order to estimate the values of the renormalisation constants we
have used the following "boosted" coupling constant $\alpha_S^V$
\cite{Lepage}
\begin{equation}
\alpha_s^V=\frac{1}{\<\frac{1}{3}\mbox{Tr } U_P\>}\alpha_s^{\mbox{ latt}}
\simeq 1.68\ \alpha_s^{\mbox{ latt}}
\ \ (\mbox{at} \,\, \beta=6.0)\, .
\end{equation}
as our expansion parameter and refer to the result as corresponding
to Boosted Perturbation Theory (BPT). We also present the values obtained
using the bare coupling $\alpha_s^{\mbox{ latt}}$, and refer to these
results as coming from standard perturbation theory (SPT).
\section{Numerical results}
\label{sec:numerical}
\begin{figure}
\vspace{9pt}
\begin{center}\setlength{\unitlength}{1mm}
\begin{picture}(160,100)
\put(10,-35)
{\special{PSfile=z4f.ps}}
\end{picture}
\end{center}
\caption{\it Non-perturbative renormalisation constants of the operator
$O_+$ as a function of $\mu^2 a^2$: (a)
the overall renormalisation constant $Z_+$; (b)--(d)
the mixing coefficients $Z_i,\ i=1,2,3$. We also report the
perturbative evaluation: the dashed curve is from BPT, while the solid
curve is from SPT.}
\label{fig:Z's}
\end{figure}
In this section, we give the numerical results of our calculation. The
simulation has been performed by generating 36 independent gluon-field
configurations, on a $16^3\times 32$ lattice, at $\beta=6.0$. The
errors have been obtained with the jacknife method, by decimating three
configurations at a time. The SW-Clover quark propagators have been
computed at a single value of the quark mass $(m_qa\simeq 0.07)$,
corresponding to the hopping parameter $\kappa=0.1425$. The quark
Green functions have been computed in the lattice Landau gauge,
defined by minimizing the functional
\begin{equation}
\mbox{Tr } \left[ \sum_{\mu=1}^4(U_{\mu}(x)+U_{\mu}^{\dag}(x))\right].
\end{equation}
Possible effects from Gribov copies have not been studied. For more
details, see ref. \cite{paciello}.
In fig. \ref{fig:Z's}, the renormalisation constants, obtained by
using the prescription described in sections \ref{sec:strategy} and
\ref{sec:mixing}, are given as a function of the renormalisation scale
$\mu^2 a^2$. We hope to find an interval of values of $\mu^2 a^2$,
large enough to avoid significant non-perturbative effects and small
enough to avoid large discretization errors. In ref. \cite{NP}, the
existence of such a ``window" in $\mu^2 a^2$ was investigated by
comparing the renormalisation constants of two-quark operators
computed in perturbation theory with the corresponding
non-perturbative determinations on quark states, and with the results
obtained by using the Ward identity method \cite{wi,wiukqcd,mm}. In
most of the cases a range of acceptable values was found in the
interval $0.8$--$0.9 \le \mu^2 a^2 \le 1.5$--$2.0$. At smaller values
of $\mu^2 a^2$, in particular in the case of the axial current and of
the pseudoscalar density (probably because of the presence of a
pseudo-Goldstone boson contribution), the non-perturbative corrections
were found to be large. For this reason, it is difficult to determine
the renormalisation constant of the axial current in this way. Only at
values of $\mu^2 a^2$ larger than $1.5$--$2.0$, a surprisingly large
value in our opinion, did discretization errors become clearly visible.
They were signalled by the fact that the renormalisation constants
computed at the same values of $\mu^2 a^2$, but with inequivalent
components of the momentum $p$ (e.g. $p\equiv 2 \pi/16 a (4,4,0,2)$ and
$p\equiv 2 \pi/16 a (0,0,0,6)$) were found to be different \cite{NP}.
This was interpreted as a signal of the breaking of the Lorentz
symmetry due to lattice artefacts, see also ref. \cite{nico}.
\begin{table}
\centering
\begin{tabular}{|c|c|c|c|c|}
\hline
$\mu^2 a^2$ &$Z_+$ & $Z_1$ & $Z_2$& $Z_3$ \\
\hline \hline
$0.46$ & $0.91 \pm 0.05$ & $0.08 \pm 0.14$ & $0.34 \pm 0.03$ & $0.34 \pm 0.07$
\\
$0.66$ & $0.84 \pm 0.05$ & $0.07 \pm 0.09$ & $0.33 \pm 0.03$ & $0.34 \pm 0.06$
\\
$0.81$ & $0.83 \pm 0.04$ & $0.11 \pm 0.07$ & $0.31 \pm 0.03$ & $0.28 \pm 0.05$
\\
$0.96$ & $0.84 \pm 0.03$ & $0.14 \pm 0.07$ & $0.30 \pm 0.02$ & $0.24 \pm 0.04$
\\
$1.27$ & $0.80 \pm 0.04$ & $0.17 \pm 0.05$ & $0.29 \pm 0.02$ & $0.21 \pm 0.03$
\\
$1.54$ & $0.82 \pm 0.02$ & $0.19 \pm 0.04$ & $0.27 \pm 0.02$ & $0.16 \pm 0.03$
\\
$1.89$ & $0.83 \pm 0.03$ & $0.22 \pm 0.05$ & $0.30 \pm 0.02$ & $0.18 \pm 0.03$
\\
$2.47$ & $0.85 \pm 0.02$ & $0.22 \pm 0.06$ & $0.33 \pm 0.02$ & $0.23 \pm 0.03$
\\ \hline
SPT & $0.91$ & $0.12$ & $0.12$ & $0.12$
\\
BPT & $0.84$ & $0.21$ & $0.21$ & $0.21$
\\ \hline
\end{tabular}
\caption{\it{Values of $Z_+$ and $Z_i$ ($i=1,2,3$) for several
renormalisation scales $\mu^2 a^2$. We also give the results
obtained at $\mu^2 a^2=1$, by using ``standard" perturbation theory
(SPT) and ``boosted" perturbation theory (BPT), using an effective
coupling $\alpha_s^V=1.68 \, \alpha_s^{{\rm latt}}$.}}
\label{tab:examples}
\end{table}
For the four-fermion operators considered in this paper, we do not
have the possibility of checking the results for the renormalisation
constants by the use of Ward identities, but it appears that a similar
situation may also occur in this case. In the region of momenta $\mu^2
a^2 \ge 0.96$, the renormalisation constants are determined with a
relatively small error (the worse case being the error of $Z_1$ which
is about $50 \%$ at $\mu^2 a^2 = 0.96$) and the dependence on the
scale is relatively weak, as can be seen from table
\ref{tab:examples} and fig. \ref{fig:Z's}. We notice that the values
of $Z_+$ and $Z_2$ have small statistical errors even at scales
smaller than $0.96$ ($\sim 10 \%$ in the worst case),
that $Z_3$ has relative errors in the range of about 15-20\%,
and that $Z_1$ has the largest relative error at all the scales
considered in table \ref{tab:examples}. As for the scale dependence,
$Z_2$ is quite stable as a function of $\mu^2 a^2$, while both $Z_1$,
which suffers from the largest statistical uncertainty, and $Z_3$ do
not exhibit a very clear plateau, as can be seen in fig. \ref{fig:Z's}.
In particular the value of $Z_1$ seems to increase
with the scale. With the present statistics, we cannot determine
whether the variation of $Z_1$ and $Z_3$ with $\mu^2 a^2$ is real or
due to statistical fluctuations. Fortunately, as we will see below, the
largest correction to the chiral behaviour comes from the operator
$O_2$, corresponding to $Z_2$, which is very well determined.
Hence the chiral behaviour of the operator $O_+$ is
stable with respect to the uncertainties above.
In order to investigate the effects of the non-perturbative
corrections, we have combined our results with the computation of the
lattice matrix elements of the four-fermion operators $O_0$-$O_3$
(\ref{eq:O_+})--(\ref{eq:O_+^SPT}) performed in ref. \cite{Crisafulli},
where a more detailed discussion of the numerical aspects can be found.
Here we limit ourselves to a qualitative discussion of the results.
In fig. \ref{fig:chiral}, we show the chiral behaviour of
$\< O_+ \>=\<\bar K^0|
O^{\Delta S=2} | K^0\>_{\mbox{ latt}}/\langle P_5 \rangle^2$
with the meson at rest
as a function of $X=8/3 f_K^2 M_K^2 /\langle P_5 \rangle^2$.
$\langle P_5 \rangle^2=\vert\langle 0 \vert \bar s \gamma_5
d \vert K^0 \rangle \vert^2$ is the squared
matrix element of the pseudoscalar density between the meson and the
vacuum. The variables $\< O_+ \>$ and $X$
are particularly
convenient since they can be obtained from suitable
two- and three-point correlation
functions without any fitting procedure \cite{Gavela}.
An analysis of the different
contributions to the matrix element of the renormalised operator shows
that the largest correction comes from the operator $O_2$, eq.\
(\ref{eq:O_+^VA}), whose constant $Z_2$ is well determined. This
contribution is much larger than that coming from $O_1$ (and
larger than that from $O_3$) \footnote{Notice that $O_1$ has
the smallest colour factor.}.
\begin{table}
\centering
\begin{tabular}{|c|c|c|c|c|}
\hline
$\mu^2 a^2$ & $\alpha$ & $\beta$& $\gamma$ \\ \hline \hline
$0.46$ & $ 0.022(16) $ & $0.23(19) $ & $ 0.78(13) $\\
$0.66$ & $ 0.013(15) $ & $0.20(17) $ & $ 0.70(12) $\\
$0.81$ & $ 0.012(13) $ & $0.21(17) $ & $ 0.69(12) $\\
$0.96$ & $ 0.017(13) $ & $0.21(17) $ & $ 0.70(12) $\\
$1.27$ & $ 0.015(13) $ & $0.21(16) $ & $ 0.66(11) $\\
$1.54$ & $ 0.018(13) $ & $0.22(16) $ & $ 0.67(12) $\\
$1.89$ & $ 0.023(13) $ & $0.22(16) $ & $ 0.69(12) $\\
$2.47$ & $ 0.022(14) $ & $0.23(17) $ & $ 0.72(12) $\\
\hline
SPT & $ -0.067(12)$ & $0.17(15)$ & $0.62(11)$ \\
BPT & $ -0.054(12)$ & $0.17(15)$ & $0.62(11)$ \\
\hline
\end{tabular}
\caption{\it{ Values of the coefficients
$\alpha$, $\beta$ and $\gamma$ obtained from a linear fit of
$\< O_+ \>=\<\bar K^0| O^{\Delta S=2} | K^0\>_{\mbox{ latt}}/\< P_5\>^2$.
The results refer to operators
renormalised at different scales $\mu^2 a^2$. Values of the same parameters
in standard perturbation theory
and in boosted perturbation theory are also reported.}}
\label{abg}
\end{table}
Thus, the uncertainty in $Z_1$ (and partly $Z_3$) has no significant
consequences for the value of the kaon matrix element (\ref{eq:B_K}).
Indeed, as can be seen from table \ref{abg}, in passing from
$\mu^2 a^2=0.66$ to $\mu^2a^2=2.47$, the central value of $Z_1$
increases by a factor of 3, but there are no large variations in the
values of $\alpha, \beta$ and $\gamma$, eq.\ (\ref{eq:B_K_latt}), i.e.
the ``physical" results depend rather weakly on the scale.
The use of the non-perturbative renormalisation constants leaves the
values of $\beta$ and $\gamma$ almost unchanged compared to those
obtained by using the constants computed in one-loop perturbation
theory: $\beta \sim 0.2$, and within the errors is compatible with zero
in both cases, and $\gamma$ is about $15 \%$ larger in the
non-perturbative case. In contrast, $\alpha$ changes sign and its
absolute value is reduced by about a factor of three in the
non-perturbative case, and becomes compatible with zero. This happens
for any choice of $\mu^2 a^2$ between $0.46$ and $2.47$, cf. tab.\
\ref{tab:examples}.
The variation of the $Z_i$'s in the interval of $\mu^2 a^2$ considered
in tables \ref{tab:examples} and \ref{abg} is representative of the
variation allowed by the statistical errors. Since the $Z_i$'s and the
matrix elements of the four-fermion operators have been computed on
different sets of configurations, this is the best test of the
stability of the results which can be done at present. A more definite
conclusion will be reached by computing the renormalisation constants
and the matrix elements on the same set of configurations.
Since $\alpha$ should vanish in the continuum limit, we conclude that
the use of the non-perturbative renormalisation constants improves the
chiral behaviour for a large range of values of the renormalisation
scale.
This is also illustrated in fig. \ref{fig:chiral}
\cite{Crisafulli}, obtained for $\mu^2 a^2=0.96$.
More details will be given in ref.\ \cite{Crisafulli}.
\begin{figure}
\vspace{9pt}
\begin{center}\setlength{\unitlength}{1mm}
\begin{picture}(160,100)
\put(-10,-60){\special{PSfile=chiral.ps}}
\end{picture}
\end{center}
\caption{\it
Chiral behaviour of $\<O_+\>$
as a function of $X$ (see text). We give the
matrix elements of the operator renormalised in standard
perturbation theory ($\Diamond$), boosted perturbation theory ($\Box$)
and non-perturbatively (\protect\mbox{\begin{picture}(2,2)(0,0)\put(1,1){\circle{2}}\end{picture}}).} \label{fig:chiral}
\end{figure}
\section{Conclusions}
\label{sec:conclusion}
We have applied the non-perturbative renormalisation method proposed in
ref. \cite{NP} to the $\Delta S=2$ operator given in eq.\
(\ref{eq:O_DS=2}). Since, on the lattice, this operator mixes with
other dimension-six operators of different chirality, we have
illustrated a projection method for the determination of the mixing
coefficients. The overall renormalisation constant of the subtracted
operator has then been obtained as in the case of any other
multiplicatively renormalisable operator.
In this exploratory study we have computed the subtraction constants
with limited statistical precision (36 configurations on a $16^3\times
32$ lattice at $\beta=6.0$ using the improved SW-Clover action). The
results in fig.\ref{fig:Z's} and table \ref{tab:examples} are very
encouraging, and motivate us to repeat the calculation with larger
statistics and at different values of $\beta$ ($\beta=6.2$ and $6.4$),
and to extend it to the operators relevant for $\Delta I=1/2$
transitions and to the penguin operators which control CP-violation in
kaon systems. Even with our limited statistical precision, our results
indicate that the chiral behaviour of the $K^0$--$\bar K^0$ matrix
element of $O^{\Delta S=2}(\mu)$ is improved significantly by the use
of the subtraction constants which were determined non-perturbatively.
This supports our view that, by combining the improvement of the action
{\em \`a la} Symanzik, which reduces $O(a)$ effects, with the
non-perturbative method of ref. \cite{NP}, which reduces higher-order
effects in the mixing coefficients, it is possible to achieve an
accurate determination of the physical weak amplitudes using
Wilson-like fermions.
\section*{Acknowledgements}
We warmly thank the members of the APE collaboration for the use of
their results before publication. M.T. thanks the Physics Department of
Southampton University for their kind hospitality during the
completion of this work. We acknowledge the partial support by the EC
contract CHRX-CT92-0051 and by M.U.R.S.T., Italy. C.T.S. acknowledges
the Particle Physics and Astronomy Research Council for its support
through the award of a Senior Fellowship. We also acknowledge the
Computer Centre of CINECA (Bologna, Italy), where these calculations
were performed, and thank their staff for their precious help.
|
\section{Introduction and Results}
It has become widely accepted that most of our Universe is made of cold dark
matter particles. Big bang cosmology implies that these particles have
interactions of order the weak scale, i.e.\ they are WIMPs\cite{Seckel}. In the
early Universe WIMPs are in equilibrium with photons. When the Universe cools
to temperatures well below the mass $m_\chi$ of the WIMP their density is
Boltzmann-suppressed as $\exp(-m_\chi/T)$ and would, today, be exponentially
small if it were not for the expansion of the Universe. At some point, as a
result of this expansion, WIMPs drop out of equilibrium with other particles
and a relic abundance persists. The mechanism is analogous to nucleosynthesis
where the density of helium and other elements is determined by competition
between the rate of nuclear reactions and the expansion of the Universe.
For WIMPs to make up a large fraction of the Universe today, i.e.\ a large
fraction of $\Omega$, their annihilation cross section has to be ``just
right''. The annihilation cross section can be dimensionally written as $
\alpha^2/m_\chi^2$, where $\alpha$ is the fine-structure constant. It then
follows that
\begin{equation}
\Omega\propto1/\sigma\propto m_\chi^2 \,. \label{omega}
\end{equation}
The critical point is that for $\Omega\simeq1$ we find that $m_\chi\simeq m_W$,
the mass of the weak intermediate boson. There is a deep connection between
critical cosmological density and the weak scale. Weakly interacting particles
which constitute the bulk of the mass of the Universe remain to be discovered.
When our galaxy was formed the cold dark matter inevitably clustered with the
luminous matter to form a sizeable fraction of the
\begin{equation}
\rho_{\chi}=0.4\rm~GeV/cm^3 \label{density}
\end{equation}
galactic matter density implied by observed rotation curves. Unlike the
baryons, the dissipationless WIMPs fill the galactic halo which is believed to
be an isothermal{\parfillskip0pt\par\noindent} sphere of WIMPs with average
velocity
\begin{equation}
v_{\chi}=300\rm\ km/sec \,. \label{velocity}
\end{equation}
In summary, we know everything about these particles (except whether they
really exist!). We know that their mass is of order of the weak boson mass
with:
\begin{equation}
\mbox{tens of GeV} < m_{\chi} < \rm several\ TeV \,. \label{GT}
\end{equation}
Lower masses are excluded by accelerator and (in)direct searches while masses
beyond several TeV are excluded by cosmological considerations. We know that
WIMPs interact weakly. We also know their density and average velocity in our
Galaxy given the assumption that they constitute the dominant component of the
density of our galactic halo as measured by rotation curves.
Two general techniques, referred to as direct (D) and indirect (ID), are
pursued to demonstrate the existence of WIMPs\cite{Seckel}. In direct detectors
one observes the energy deposited when WIMPs elastically scatter off nuclei.
The indirect method infers the existence of WIMPs from observation of their
annihilation products. WIMPs will annihilate into neutrinos; massive WIMPs will
annihilate into high-energy neutrinos which can be detected in high-energy
neutrino telescopes. Throughout this paper we will assume that such neutrinos
are detected in a generic Cherenkov detector which measures the direction and,
to some extent, the energy of a secondary muon produced by a neutrino of WIMP
origin in or near the instrument\cite{Gaisser}. It can also detect the showers
initiated by electron-neutrinos.
The indirect detection is greatly facilitated by the fact that the sun
represents a dense and nearby source of accumulated cold dark matter
particles\cite{Drees}. Galactic WIMPs, scattering off nuclei in the sun, lose
energy. They may fall below escape velocity and be gravitationally trapped.
Trapped WIMPs eventually come to equilibrium temperature and accumulate near
the center of the sun. While the WIMP density builds up, their annihilation
rate into lighter particles increases until equilibrium is achieved where the
annihilation rate equals half of the capture rate. The sun has thus become a
reservoir of WIMPs which we expect to annihilate mostly into heavy quarks and,
for the heavier WIMPs, into weak bosons. The leptonic decays of the heavy quark
and weak boson annihilation products turn the sun into a source of high-energy
neutrinos with energies in the GeV to TeV range.
The performance of future detectors is determined by the rate of elastic
scattering of WIMPs in a low-background, germanium detector and, for the
indirect method, by the flux of solar neutrinos of WIMP origin. Both are a
function of WIMP mass and of their elastic cross section on nucleons. In
standard cosmology WIMP capture and annihilation interactions are weak, and we
will suggest that, given this constraint, dimensional analysis is sufficient to
compute the scattering rates in germanium detectors as well as the neutrino
flux from the measured WIMP density in our galactic halo. We will derive and
compare rates for direct and indirect detection of weakly interacting particles
with mass $m_\chi \simeq m_W$ assuming
\begin{enumerate}
\advance\itemsep by -0.05in
\item
that WIMPs represent the major fraction of the measured halo density. Their
flux is
\begin{equation}
\phi_\chi = n_\chi v_\chi = {0.4\over m_\chi} \, {\rm {GeV\over cm^3} \
3\times10^6 {cm\over s} } = {1.2\times10^7\over m_{\chi\rm\,GeV}} \,\rm
cm^{-2} s^{-1} \;,
\label{phi chi}
\end{equation}
where $m_{\chi\rm\,GeV} \equiv (m_\chi/$1~GeV) is in GeV units.
\item
a WIMP-nucleon interaction cross section based on dimensional analysis
\begin{equation}
\sigma(\chi N) = \left(G_F m_N^2\right)^2 {1\over m_W^2} \equiv \sigma_{\rm DA}
= 6\times10^{-42}\rm\,cm^2 \;,
\label{sigma chi N}
\end{equation}
\item
that WIMPs annihilate 10\% of the time in neutrinos (this is just the leptonic
branching ratio of the final state particles in the dominant annihilation
channels $\chi\bar\chi \to W^+W^-$ or $Q\bar Q$, where $Q$ is a heavy quark).
\end{enumerate}
Clearly the cross section for the interaction of WIMPs with matter is
uncertain. Arguments can be invoked to raise or decrease it. Important points
are that i) our choice represents a typical intermediate value, ii) all our
results for event rates scale linearly in the cross section and can be easily
reinterpreted, and iii) the comparison of direct and indirect event rates is
independent of the choice.
We present a simple and totally transparent analysis in which the event rates
of detectors are derived from the above assumptions. It finesses all detailed
dynamics and gives answers that are sufficiently accurate considering that the
mass of the particle has not been pinned down. We will find that the event rate
in a direct detector is proportional to the WIMP cross section and flux and the
density of targets $m_N^{-1}$, i.e.
%
\begin{equation}
{dN_{\rm D} \over dM} = {1\over m_N} \phi_\chi \sigma_{\rm DA} N(A_D) =
{1.4\over m_{\chi\rm\,GeV}} \rm\ (kg)^{-1} \, (year)^{-1} \nonumber\\,
\label{direct}
\end{equation}
where ${dN_{\rm D}\over dM}$ represents the number of direct events per unit of
target mass. $N(A_D$) represents the coherent enhancement factor for a nuclear
target of atomic number $A_D$, e.g.\ 76 for Germanium,
\begin{equation}
N(A) \equiv A^3 \left[ 1 + {m_\chi\over m_N} \over A + {m_\chi\over m_N}
\right]^2 \;.
\label{nuclear}
\end{equation}
The rates for indirect detection are
\begin{equation}
dN_{\rm ID}/ dA \simeq \left\{ 1.8\times10^{-2}m_{\chi\rm\,GeV} \right\}
\left\{ \rho^{\vphantom0}(A_{ID}) N(A_{ID})\right\} \left\{ 1+1.9\times10^{-4}
m_{\chi\rm\,GeV} \right\}^{-7} \;,
\label{indirect}
\end{equation}
where $dN_{\rm ID} / dA$, in units of $\rm\ (10^4\,m^2)^{-1} (year)^{-1}$,
represents the number of events from the sun per unit area $A$ detected by a
neutrino telescope. The factor \{$\rho N$\} should be summed over all elements
in the sun. Because of additional nuclear form factor effects which are
neglected in Eq.~\ref{indirect} it is adequate to consider oxygen with a solar
abundance of $\rho =1.1$~\% and $A_{ID} = 16$ as a ``typical'' element. The
observed average muon energy should be in the range $1/4 \sim 1/6
m_{\chi\rm\,GeV}$.
The above parametrizations readily lead to the conclusion that the direct
method is superior if the WIMP interacts coherently on nuclei (which has been
assumed for Eqs.~\ref{direct}--\ref{indirect}) and, if its mass is lower or
comparable to the weak boson mass $m_W$. We will show that in all other cases,
i.e.\ for relatively heavy WIMPs and for all WIMPs interacting incoherently,
the indirect method is competitive or superior, but it is, of course, held
hostage to the successful deployment of high energy neutrino telescopes with
effective area in the $\sim10^4$--$10^6$~m$^2$ range and with appropriately low
threshold. Especially for heavier WIMPs the indirect technique is powerful
because underground high energy neutrino detectors have been optimized to be
sensitive in the energy region where the neutrino interaction cross section and
the range of the muon are large. A kilometer-size detector probes WIMP masses
up to the TeV-range, beyond which they are excluded by cosmological
considerations.
For high energy neutrinos the muon and neutrino are aligned, with good angular
resolution, along a direction pointing back to the sun. The number of
background events of atmospheric neutrino origin in the pixel containing the
signal will be small. The angular spread of secondary muons from neutrinos
coming from the direction of the sun is well described by the
relation\cite{Gaisser} $\sim 1.2^\circ \Big/ \sqrt{E_\mu(\rm TeV)}$.
Measurement of muon energy, which may be only up to order of magnitude accuracy
in some experiments, can be used to infer the WIMP mass from the angular spread
of the signal. The spread contains information on the neutrino energy and,
therefore, the WIMP mass. More realistically, measurement of the muon energy
can be used to reduce the search window around the sun, resulting in a reduced
background.
\looseness=-1
Before proceeding, we comment on our ansatz for the elastic WIMP-nucleon
scattering cross section. The simplest dimensional analysis implies that the
cross section is $G_F^2 m_N^2$. This correctly describes the $Z$-exchange
diagram of Fig~1a, which is of the form
\begin{equation}
\sigma\ \sim\ G_F^2 {m_N^2m_\chi^2\over (m_N+m_\chi)^2} \;.
\label{sigd}
\end{equation}
\vskip-.1in
\begin{figure}[h]
\centering
\epsfxsize=3.25in\hspace{0in}\epsffile{fig1.eps}
\medskip
\parbox{6in}{\footnotesize\baselineskip13pt Fig.~1. Examples of (a) incoherent
and (b)~coherent WIMP-nucleon interactions. In (b) the gluon is a constituent
of the target nucleon and $Q$ is a heavy quark.}
\end{figure}
For coherent interactions, which we will emphasize throughout this paper, there
is an additional suppression factor associated with the exchange of the Higgs
particle with a mass of order of the weak boson mass; see Fig~1b. In the
specific diagram shown the Higgs interacts with the heavy quarks in the gluon
condensate associated with the nucleon target. It is of the form
\begin{equation}
\sigma \sim G_F g_H^2 {m_N^2 m_\chi^2\over(m_N+m_\chi)^2} {1\over m_W^2} \;,
\label{sigid}
\end{equation}
where $g_H \sim \sqrt{G_F}\, m_N$ describes the condensate. Conservatively, we
will use the suppressed WIMP interaction cross section which is appropriate for
coherent scattering.
\section{Derivation of Detection rates}
For the case of direct detection the structure of Eq.~\ref{direct} is
transparent\cite{Seckel}. For indirect detection the number of solar neutrinos
of WIMP origin can be calculated in 5 easy steps by determining:
\begin{itemize}
\advance\itemsep by -0.05in
\item
the capture cross section in the sun, which is given by the product of the
number of target nucleons in the sun and the elastic scattering cross section
\begin{equation}
\sigma_\odot = f \left[ 1.2\times10^{57} \right] \sigma_{\rm DA} \;.
\label{sigma sun}
\end{equation}
This includes a focussing factor $f$ given, as usual, by the ratio of kinetic
and potential energy of the WIMP near the sun. It enhances the capture rate by
a factor 10.
\item
the WIMP flux from the sun which is given by
\begin{equation}
\phi_\odot = \phi_\chi \sigma_\odot / 4\pi d^2 \;,
\label{phi sun}
\end{equation}
where $d=1\rm~a.u. = 1.5\times10^{13}\,cm$.
\item
the actual neutrino flux, which is obtained after inclusion of the branching
ratio. From (\ref{phi chi}),(\ref{sigma chi N}) and (\ref{sigma sun}),(\ref{phi
sun})
\begin{equation}
\phi_\nu = 10^{-1} \times \phi_\odot
= {3\times10^{-5}\over m_{\chi\rm\,GeV}}\rm \, cm^{-2} \, s^{-1} \;.
\end{equation}
\item
the probability to detect the neutrino\cite{Gaisser}, which is proportional to
\begin{eqnarray}
&&P = \rho\sigma_\nu R_\mu,\rm\ with\nonumber\\
&&\rho = \mbox{Avagadro\,\#} = 6\times10^{23}\nonumber\\
&&\sigma_\nu = \mbox{neutrino interaction cross section} = 0.5\times10^{-38}\
E_\nu\rm (GeV)\ cm^{2}\nonumber\\
&&R_\mu = \mbox{muon range} = 500{\rm\ cm}\ E_\mu\rm(GeV)\, \nonumber\\
\noalign{\vskip2pt}
\rm or&&P = 2\times 10^{-13} \, m_{\chi\rm\,GeV}^2
\end{eqnarray}
Here we assumed the kinematics of the decay chain
\begin{eqnarray}
\chi\bar\chi &\to& W^+W^- \nonumber \\ \noalign{\vskip-1ex}
& & \hspace{2em} \raise1ex\hbox{$\vert$}\!{\rightarrow}
\mu\nu_\mu \nonumber
\end{eqnarray}
with $E_\nu = {1\over2}m_\chi$ (this would be ${1\over3}m_\chi$ for $Q$ decay)
and $E_\mu = {1\over2}E_\nu = {m_\chi\over 4}$.
\item
finally, $dN_{\rm ID} / dA = \phi_{\nu} P = 1.8\times10^{-6} \,
m_{\chi\rm\,GeV} \, \rm\ (year)^{-1} \, (m^2)^{-1}$
\stepcounter{equation}\hfill(\theequation)\break
where $dN_{\rm ID} / dA$ represents the number of events from the sun per unit
area (m$^2$) detected by a neutrino telescope.
\end{itemize}
\noindent
We can now summarize our results so far by comparing a $10^4$~m$^2$ neutrino
detector, an area typical of the instruments now being deployed, with a
kilogram of hydrogen:
\begin{eqnarray}
dN_{\rm ID}/ dA &=& 1.8\times10^{-2} m_{\chi\rm\,GeV} \rm\ (10^4\,m^2)^{-1}
(year)^{-1} \nonumber\\
dN_{\rm D}/ dM &=& {1.4\over m_{\chi\rm\,GeV}} \rm\ (kg)^{-1} \, (year)^{-1}
\nonumber\\
{dN_{\rm D}/dM\over dN_{\rm ID}/dA} \left(10^4\rm\,m^2\over\rm kg\right)
&=& {7.8\times10^1\over m_{\chi\rm\,GeV}^2} \label{D/ID}
\end{eqnarray}
Direct detection is superior only in the mass range $m_\chi<10$~GeV, but this
region is, arguably, ruled out by previous searches. Indirect detection is the
preferred technique. This straightforward conclusion may, however, be
invalidated when WIMPs interact coherently with nuclei and targets other than
hydrogen are considered. We discuss this next.
\section{Coherent Nuclear Enhancements}
For WIMPs interacting coherently with nuclei in the detector or in the sun, the
nuclear dependence of the event rates resides in
\begin{itemize}
\advance\itemsep by -0.1in
\item
the target density factor $m_N^{-1}$ in Eq.~\ref{direct} which is replaced by
$(Am_N)^{-1}$,
\item
the coherent enhancement factor ``$A^2$",
\item
the nuclear dependence of the cross sections of Eqs.~\ref{sigd},\ref{sigid}
which is obtained by the substitution
\begin{eqnarray}
\noalign{\qquad\underline{\rm incoherent}}
\sigma &\sim& G_F^2 {m_N^2m_\chi^2\over (m_N+m_\chi)^2} \to G_F^2 {(Am_N)^2
m_\chi^2\over (Am_N+m_\chi)^2} \nonumber\\
\noalign{\qquad\underline{\rm coherent}}
\sigma &\sim& G_F^2 g_H^2 {m_N^2 m_\chi^2\over(m_N+m_\chi)^2} {1\over m_W^2}
\to G_F\left(g_H\over m_W\right)^2 {(Am_N)^2 m_\chi^2\over (Am_N+m_\chi)^2} A^2
\nonumber
\end{eqnarray}
The coherent enhancement factor for a nucleus $A$, including a factor $A^{-1}$
for the target density, is therefore given by
\begin{equation}
{1\over A} {A^2 (Am_N)^2 m_\chi^2 \over (Am_N+m_\chi)^2 } \,
{(m_N+m_\chi)^2\over m_N^2 m_\chi^2}
= A^3 {(m_N+m_\chi)^2\over (Am_N+m_\chi)^2}
= A^3\left[ 1+{m_\chi\over m_N} \over A+{m_\chi\over m_N} \right]^2 .
\end{equation}
\end{itemize}
This yields Eq.~\ref{nuclear}.
\section{Event Rates for WIMPs with Coherent Interactions}
Our simple evaluations, made so far, overestimate the indirect rates for very
heavy WIMPS because high energy neutrinos, created by annihilation near the
core, may be absorbed in the sun. Absorption is stronger for neutrinos and,
therefore, mostly antineutrinos form the signature for very heavy WIMPS. The
probability that an antineutrino escapes without absorption is well
parametrized by $(1+ 3.8 \times 10^{-4} E_{\nu})^{-7}$, where $E_{\nu} \simeq
m_{\chi}/2$. This yields our final result of Eq.~\ref{indirect}.
The relative merits of the two methods are illustrated in the following table,
which is obtained from Eqs.~\ref{direct}--\ref{indirect} and establishes that a
kilogram of germanium and a $10^4$~m$^2$ are competitive.
\vskip-.3in
\begin{table}[h]
\tcaption{Event rates and signal to background $(N/B)$.}
\centering\unskip\smallskip
\tabcolsep=1.5em
\begin{tabular}{c|c@{\quad}cc@{\quad}c}
\hline
$m_\chi$ (GeV)&\multicolumn{2}{c}{Direct (/kg/year)}&
\multicolumn{2}{c}{Indirect (/$10^4$\,m$^2$/year)}\\
\hline
& events& $N/B$& events& $N/B$\\
50 & $2.2\times10^3$& 7& $2.3\times10^1$& $\simeq\,1$\\
500 & $1.1\times10^3$& 7& $2\times10^2$& $\simeq\,10^2$\\
2000 & $2.9\times10^2$& 1& $1.7\times10^2$& $\simeq\,10^4$\\
\hline
\end{tabular}
\end{table}
At the lower energy the event rates for the indirect method are underestimated
because also the Earth is a source of neutrinos of WIMP origin.
We conclude that the direct method yields more events for the lower masses,
even when compared to a $10^6$~m$^2$ detector. As expected, the indirect method
is competitive for heavier WIMPs with a detection rate growing like $E_\nu^2$
or $m_\chi^2$. A $10^5$~m$^2$ instrument covers the full WIMP mass range, even
if the WIMPs do not coherently interact with nuclei in the sun. These
conclusions are reinforced after considering the signal-to-noise for both
techniques which we discuss next.
\section{Backgrounds}
\noindent
\underline{Indirect Background}.
For the indirect detection the background event rate is determined by the flux
of atmospheric neutrinos in the detector coming from a pixel around the
sun\cite{Gaisser}. The number of events in a $10^4$~m$^2$ detector is $\sim
10^2/E_\mu$(TeV) and the pixel size is determined by the angle between muon and
neutrino $\sim 1.2^\circ \Big/ \sqrt{E_\mu(\rm TeV)}$. Using the kinematics
$E_\mu \simeq m_\chi/4$ we obtain
\[
B_{\rm ID} = { 10^2/E_\mu({\rm TeV}) \over 2\pi \Big/
\left[ 1.2^\circ {\pi\over 180^\circ} \over \sqrt{E_\mu(\rm TeV)} \right]^2}
= {1.1\times10^5\over m_{\chi\rm\,GeV}^2} \mbox{ per 10$^4$\,m$^2$ per year}
\]
This is only valid for large $m_\chi$, i.e.\ for $E_\mu \cong m_\chi/4
>100$~GeV. Without this approximation we obtain
\vskip-.11in
\begin{table}[h]
\centering\unskip
\tabcolsep=1.25em
\begin{tabular}{c|ccc}
\hline
&\small \# bkgd. events&\small \# pixels of solar&\small bkgd. events\\
\noalign{\vskip-.85ex}
&\small in 10$^4$\,m$^2$&\small size in $2\pi$&\small per 10$^4$\,m$^2$\\
\noalign{\vskip-.85ex}
$E_\mu$(GeV) &\small in $2\pi$ &\small &\small per pixel, per year \\
\hline
10& 3200& 140& 23\\
100& 1060& $1.4\times10^3$& 0.8\\
1000& 110& $1.4\times10^4$& $8\times10^{-3}$\\
\hline
\end{tabular}
\end{table}
\noindent
For large $m_\chi$ the signal to background ratio is
\[
\left(N\over B\right)_{\rm ID} \equiv {dN_{\rm ID}/dA\over dB_{\rm ID}/dA}
\simeq
7.2\times10^{-6} m_{\chi\rm\,GeV}^3
\]
Clearly, the extremely optimistic predictions for signal-to-noise are unlikely
to survive the realities of experimental physics. One expects, typically, to
measure muon energy only to order-of-magnitude accuracy in the initial
experiments. The energy of showers initiated by electron neutrinos should be
determined to a factor 2. It is not excluded that future, dedicated experiments
may do better. The conclusion that high energy muons pointing at the sun
represents a superb signature, is unlikely to be invalidated.
\smallskip
\noindent
\underline{Direct Background}: about 300 events per year per
kg\cite{Kamionkowski}. Signal-to-noise therefore exceeds unity up to 2~TeV WIMP
mass.
These considerations were used to estimate the signal-to-noise $N/B$ in
Table~1.
\section{Dynamics?}
We emphasize that our results are representative for the specific and much
studied example where the lightest supersymmetric particle is Nature's
WIMP\cite{Drees}. Clearly dynamics, which is now defined, can alter our
conclusions, but only in ``conspiratorial" ways. Dynamics can, on the other
hand, increase rates as well, sometimes by well over an order of magnitude,
over and above the rates obtained from dimensional analysis in this paper. Our
qualitative conclusions are valid, at least in some average sense, in
supersymmetry. Our results do, in fact, closely trace the supersymmetry
prediction of reference 2 for the choice of Higgs coupling ${g_H^2 \over 4
\pi}=1$, in their notation.
We feel that the development of detectors should be guided by an analysis like
ours rather than by dynamics of theories beyond the standard model for which
there is, at present, no experimental guidance.
\bigskip
\leftline{\bf Acknowledgements}
\medskip
We thank M.~Drees, S.~Pakvasa and X.~Tata for a careful reading of the
manuscript.
This research was supported in part by the U.S.~Department of Energy under
Grant No.~DE-FG02-95ER40896 and in part by the University of Wisconsin Research
Committee with funds granted by the Wisconsin Alumni Research Foundation.
\nonumsection{References}\unskip
|
\section{Introduction}
It has been known since the pioneering work of Kirzhnits and Linde \cite{Linde}
and Dolan and Jackiw \cite{Jackiw} that, at high temperature, the thermal
populations of massive particles exert a symmetry restoring force on the
Higgs condensate, so that at sufficiently high temperaure ($T \sim m_H/g$)
the Higgs field loses its condensate and electroweak symmetry is restored.
Under these circumstances baryon number is readily violated
\cite{Rubakov,McLerran}. As temperature falls there is a symmetry breaking
phase transition, which is first order for small $m_H$. Presumably the
universe underwent such a phase transition shortly after the Big Bang, and
it is believed likely that the baryon asymmetry of the universe was
generated at this time.
Equilibrium thermodynamic information important to the physics of
this epoch, such as the effective potential
and the free energy of certain saddlepoint solutions,
can be computed in the Matsubara (imaginary time) formulism.
The particular calculation we will be interested in is the Sphaleron rate
in the broken phase shortly after the phase transition. This was first
investigated at tree level in \cite{Klink}, and one loop corrections have
been computed within a high temperature approximation in
\cite{Carson,Junker}. This approximation assumes that the contributions of
nonzero Matsubara frequencies can be absorbed into shifts in the parameters
of the zero Matsubara frequency modes, which then constitute a ``dimensionally
reduced'' three dimensional theory\cite{Landsman}.
The same idea forms the basis for
an extensive investigation of the effective potential and the strength
of the phase transition being carried out by Farakos, Kajantie, Laine,
Rummukainen, and Shaposhnikov (FKLRS) \cite{FKRS,FKLRS}.
Why is it necessary to include 1 loop corrections?
Normally, one loop corrections
modify tree level results by only a few percent, but there are exceptions.
The finite temperature, 1 loop correction to the Higgs mass is significant
because, although the correction
is parametrically order $g^2$, the mass is a dimensionful
parameter, and the importance of the thermal effect is enhanced by $T^2/m^2$,
which can be large. The phase transition temperature is
the point where the thermal, 1 loop effect and the tree level effect are
almost equal. Finite temperature calculations also have potential infrared
divergences from loops containing zero Matsubara frequency bosons,
and it is not clear that perturbation theory always works; the object of
the dimensional reduction program is to separate this problem from those
effects which can be treated perturbatively more reliably. Finally,
loop corrections can also be important when a coupling
is unnaturally small; for instance, the Higgs self coupling $\lambda$ recieves
a 1-loop correction which goes as $g_t^4$ ($g_t$ is the Yukawa coupling of
the top quark. In this paper we will always take $g_t=1$, corresponding to
$m_t \simeq 174$GeV). This correction does not depend on $\lambda$, and if
$\lambda$ happens to be small
then the correction can be very important. This is the case in the
standard model whith a light Higgs boson and a heavy top quark. It is
quite necessary, then, to consistently use 1 loop relationships when
describing anything directly related to $\lambda$, such as the physical
Higgs mass and the (finite and zero temperature) effective potential. In
this case, however, two loop effects are still expected to make a small
perturbative correction to the one loop effects, so the one loop calculation
should be reliable.
With this in mind, Diakonov, Polyakov, Sieber, Schaldach, and Goeke
(DPSSG) have made a direct numerical evaluation of the fermionic fluctuation
determinant and have concluded on its basis that, when the
top quark is heavy, fermion fluctuations make the Sphaleron rate very
different from the high temperature estimate \cite{DPSSG}. This brings the
dimensional reduction program into question, and bears further investigation.
Here we give a one loop, perturbative treatment of
the fermions in the presence of zero Matsubara frequency background fields.
The free energy of the fermions can naturally be expressed in terms of
operators made up of zero Matsubara frequency bosonic fields and their
derivatives. In section II we compute these, including induced masses
(dimension 2), corrections to couplings and wave function renormalizations
(dimension 4), and nonrenormalizeable dimension 6 operators. The expansion
in operator dimension appears to be very well behaved when the background
bosonic fields are slowly varying.
In section III we apply these results to a computation of the
Sphaleron energy. In section IV we use this calculation, along with 1-loop
relationships between the couplings and the physical Higgs mass and a
calculation of the bubble nucleation rate which determines the temperature
at which the phase transition occurs, to investigate the erasure of
baryons after the phase transition is complete. In the last section we
draw conclusions. Some technical details are relegated to an appendix.
\section{Integration over fermions}
We will work in SU(2) Higgs theory coupled to all the fermions of the
standard model. Neglecting hypercharge significantly simplifies the
picture whithout profoundly changing it, because the Weinberg angle is
small and because the fields which make up the Sphaleron are almost
exclusively weak isospin, and not hypercharge, fields.
(It is known that, at tree level,
including hypercharge with the physical Weinberg angle only changes the
Sphaleron energy by about 1\% \cite{Klink}). The Lagrangian, suppressing
summations on generations and colors, is
\begin{eqnarray}
{\cal L} & = & {\cal L}_{b} + {\cal L}_{f}
\nonumber \\
{\cal L}_{b} & = &
\frac{1}{4} F_{\mu \nu}^a F_{\mu \nu}^a + (D_\mu \Phi)^{\dag} D_\mu \Phi
- m_0^2 \Phi^{\dag} \Phi + \lambda (\Phi^{\dag} \Phi)^2
\nonumber \\
{\cal L}_{f} & = & \overline{\psi}_L \gamma_\mu D_\mu \psi_L
+ \overline{\psi}_R \gamma_\mu \partial_\mu \psi_R + g_t (\overline{\psi}_L
\Phi \psi_R + \overline{\psi}_R \Phi^{\dag} \psi_L ) \, .
\nonumber
\end{eqnarray}
(We use a Euclidean metric
$g_{\mu \nu} = \delta_{\mu \nu}$ and Euclidean $\gamma$ matricies which
satisfy the algebra $ \gamma_\mu \gamma_\nu + \gamma_\nu \gamma_\mu =
2 \delta_{\mu \nu}$.)
The fermion kinetic terms apply for every doublet and singlet, but the
mass term only applies for the top quark. Actually the mass term as written
is for a bottom, not top, type quark. This is for
notational simplicity only; to make the top quark massive one should
systematically replace $\Phi$ with $i \tau_2 \Phi^*$ in all that
follows; our conclusions are completely unchanged.
In the 1 loop approximation we ignore the fermions' coupling to nonzero
Matsubara frequency excitations of the bosonic fields. The path integral
over the fermions is gaussian; its contribution to the partition function
is Det$H$, and the corresponding contribution to the effective action
is $-$Tr ln$H$, where $H$, written as a matrix acting
on $[\psi_L^\alpha \; \; \psi_R]^T$, is
\begin{equation}
H = \left[ \begin{array}{cc} \gamma_\mu D^\mu_{\beta \alpha}
& g_t \Phi_\beta \\
g_t \Phi^{\dag}_\alpha & \gamma_\mu \partial^\mu \end{array} \right]
= \left[ \begin{array}{cc} \gamma_{\mu} \partial^{\mu}
\delta_{\beta \alpha} & 0 \\ 0 & \gamma_{\mu} \partial^{\mu}
\end{array} \right] + \left[ \begin{array}{cc}
\frac{ig_w}{2} \gamma_{\mu} A^{\mu}_{a} \tau^{a}_{\beta \alpha} &
g_t \Phi_{\beta} \\ g_t \Phi^{\dag}_{\alpha} & 0 \end{array} \right]
\equiv H_0 + H_I \, ,
\label{Hamiltonian}
\end{equation}
where $\alpha$ and $\beta$ are $SU(2)$ indicies.
Our analysis will be based on an expansion of $-{\rm Tr} \ln H$ in
$H_I$. To illustrate the idea of expanding the log,
consider a simplified example
in which the gauge field is everywhere zero. In this case we have a Dirac
fermion with a (spatially varying) mass, $H = \gamma^\mu \partial_\mu - m$,
where $m^2 = g_t^2 \Phi^{\dag} \Phi$.
We assume that $m^2$ is smaller than the lowest eigenvalue of
$- \partial^2$, in which case it is legitimate to expand the log.
This will generally be the case, as the lowest eigenvalue of $-\partial^2$
is set by the square of the lowest possible Matsubara frequency, which for
a fermion is $(\pi T)^2$. The log becomes
\begin{equation}
-{\rm Tr} \, \ln ( \gamma^\mu \partial_\mu - m ) =
-{\rm Tr}\, \ln \gamma^\mu \partial_\mu
- \sum_{n=1}^{\infty} \frac{(-1)^{n+1}}{n} {\rm Tr} \left(
\frac{m}{\gamma_\mu \partial_\mu} \right)^n
\end{equation}
where as usual $1/(\gamma^\mu \partial_\mu)$ is defined as $\gamma^\mu
\partial_\mu / \partial^2$. We get Feynman diagrams by the usual trick of
inserting complete sets of states and Fourier transforming. The first term
in the sum is a divergent vacuum energy and should be removed. All terms with
odd powers of $m$ vanish when we take the trace on Dirac indicies. When
the mass is position independent, the resulting terms are
\begin{equation}
\int_0^{T^{-1}} \! \! dx_0 \int d^3 x
\sum_{n=1}^{\infty} \, (-)^{n} \frac{4m^{2n}}{2n} T \sum_{k_0}
\int \frac{d^3 k}{(2\pi)^3} \frac{1}{(k^2 + k_0^2)^n} \, .
\label{massexpan}
\end{equation}
We have used the shorthand $\sum_{k_0}$ to mean a sum in which $k_0$ takes
on odd integer multiples of $\pi T$; we will use this shorthand throughout.
Also, $k^2$ means $\vec{k}^2$ and $k$ means $\sqrt{\vec{k}^2}$.
The $(-)^n$ arises because when we Fourier transform $\partial \rightarrow
ik$. The $4$ is from the Dirac trace, and $2n$ is the symmetry factor
of the diagram; we saw how it arises in the expansion of the log above.
To get the free energy density we drop the space integral.
The first two integrals in the series are ultraviolet divergent and must be
performed with some care. We have done so in the $\overline{\rm MS}$ scheme
by first performing the sum on Matsubara modes and conducting the spatial
integrals in $3-2 \epsilon$ dimensions.
The higher order terms are convergent and may be performed directly by doing
the integral over $d^3k$ first. Performing the
integrals, the free energy per unit volume is
\begin{equation}
m^2 \frac{T^2}{12} + m^4 \frac{2\gamma_E - 2 \ln \pi + \ln (\mu^2/T^2)}
{16\pi^2} + \sum_{n=3}^{\infty} (-)^n \frac{4 m^{2n}}{2n}
\frac{ (1-2^{3-2n}) \zeta(2n-3) \Gamma(n-\frac{3}{2})}{4 \pi^{2n-2} T^{2n-4}
\Gamma(\frac{1}{2}) \Gamma(n)}
\label{hiTexpan}
\end{equation}
with $\mu$ the usual $\overline{\rm MS}$ renormalization point.
This is the 1 loop fermionic contribution to the finite temperature effective
potential, to all orders in the high temperature expansion\footnote{An
integral expression for the fermionic contribution was first found by
Dolan and Jackiw \cite{Jackiw}, who also found the first two terms in the
expansion given here. What we have done is found the complete Taylor series
for the known integral expression.}.
Note that, as expected, the sum is a Taylor
series in $m^2$ with radius of convergence $(\pi T)^2$. By $m_t$ we mean
the thermal top quark mass given by $m_t^2(T) = g_t^2 \Phi^{\dag} \Phi (T)$,
not the vacuum top quark mass $m_t^2(0) = g_t^2 \Phi^{\dag} \Phi (T = 0)$.
Near the phase transition temperature the difference is quite substantial.
In the simplest approximation $\Phi^{\dag} \Phi (T) = \Phi^{\dag} \Phi(T=0)
(1-T^2/T_c^2)$, but near $T_c$ it is necessary to treat the influence of
infrared bosons more carefully. As we will see in Sec. IV, at the
temperature of interest the Higgs VEV
$\nu \equiv \sqrt{\Phi^{\dag} \Phi/2} \leq T$ and
for $g_t=1$ we find $(m/\pi T)^2 \leq 1/(2 \pi^2)$, so the convergence of
the series is excellent. Of course, well below the phase transition
temperature it is no longer true that $m_t << \pi T$ and our approximation
scheme will fail. However, at these temperatures the Sphaleron energy
is so large that the baryon erasure rate is utterly negligible, so our
approximation scheme should apply in the interesting range of temperatures.
Another way of understanding the convergence of the power series is to think
of the different Matsubara frequency contributions of the top quarks as
distinct, very massive species of a three dimensional theory, with masses
given by the Matsubara frequencies $m_M^2 = ( (2n+1) \pi T)^2$; we are
then expanding in the ratio of $g_t^2 \Phi^{\dag} \Phi$ (and eventually
in products of other infrared bosonic fields like $A_i$, and their
derivatives) to $m_M^2$. Such an expansion has been considered in a theory
with only fermions and a scalar in \cite{Baacke1,Baacke2}, where it
is also concluded that such an expansion is very accurate. An expansion
like the one used here is not justified for bosons because
symmetric boundary conditions in time make the lowest
eigenvalue of the operator $\partial^2$ is zero; or equivalently the
lowest Matsubara frequency is zero, so we cannot expand in the ratio of
a field value to this Matsubara frequency. (However we could expand all
the other Matsubara frequencies in the way described here, see \cite{FKRS}.)
For simplicity of notation, in the remainder of the paper we write
\begin{equation}
T \sum_{k_0} \int \frac{d^3k}{(2\pi)^3}
\frac{1}{(k^2 + k_0^2)^2} =
\frac{2 \gamma_E - 2 \ln \pi + \ln (\mu^2 / T^2)}{16 \pi^2} \equiv D4
\label{D4def}
\end{equation}
and
\begin{equation}
T \sum_{k_0} \int \frac{d^3k}{(2\pi)^3}
\frac{1}{(k^2 + k_0^2)^N}
= \frac{(1-2^{3-2N}) \zeta(2N-3)
\Gamma(N-\frac{3}{2})}{4\pi^{2N-2} T^{2N-4} \Gamma(\frac{1}{2}) \Gamma(N)}
\equiv D2N \, .
\end{equation}
in particular,
\begin{equation}
D6 = \frac{7 \zeta(3)}{128 \pi^4 T^2} \, .
\end{equation}
When the mass is spatially varying the first term in the
sum in Eq.(\ref{massexpan}) becomes
\begin{equation}
\int dx_0
\frac{-1}{2} \int \frac{d^3p}{(2 \pi)^3} m(p) m(-p)
T \sum_{k_0}
\int \frac{d^3 k}{(2\pi)^3}
\frac{ {\rm tr} ( \gamma^\mu \gamma^\nu ) k_\mu (p+k)_\nu}
{(k^2 + k_0^2) ((p+k)^2 + k_0^2)} \, .
\label{derivexpan}
\end{equation}
If we assume that $p^2 < k^2 + k_0^2$, which means that the mass
varies on a scale
large compared to $1/\pi T$, then we may expand the denominator in a geometric
series and extract a power series in $p^2$. We can then Fourier transform
to position space to express the result as a derivative expansion for $m$.
The resulting free energy is
\begin{equation}
\int d^3x \left( \frac{T^2}{12} m^2 + D4 \, (\vec{\nabla} m)^2
- \frac{D6}{3} (\nabla^2 m)^2
+ \frac{D8}{10} ( \vec{\nabla} \nabla^2 m)^2
\ldots \right) \, .
\label{asymptotic}
\end{equation}
In realistic cases the Fourier transform of $m$ lies primarily at small
values of $p$ but has a rapidly decaying exponential tail which goes
above $p = \pi T$. In this case the series will be asymptotic and its
reliability will depend on how rapidly the exponential tail falls off.
For the Sphaleron at $T \simeq T_c$ we expect the exponential tail of
the Fourier transform of the gauge and Higgs field configurations to
fall roughly as $\exp(-p/m_W (T))$; at $T \simeq T_c$,
$m_W(T) = g_w \nu(T) /2$ is $<< \pi T$ and the convergence
of the (asymptotic) derivative expansion should be excellent. Of course
we will check this by explicit calculation.
For the more general $H$ the lowest nonvanishing term will be
\begin{eqnarray}
\frac{1}{2} {\rm Tr} \frac{H_I}{H_0^2} H^0 \frac{H_I}{H_0^2} H_0
\Rightarrow \frac{1}{2} \int \frac{d^3 p}{(2 \pi)^3}
T \sum_{k_0} \int \frac{d^3 k}{(2 \pi)^3}
(\Phi^{\dag}_{\alpha}(p) \Phi_{\alpha}(-p))
\frac{i k_{\mu} i (k+p)_{\nu} {\rm tr}(\gamma^{\mu} \gamma^{\nu})}{(k^2+k_0^2)
( (k+p)^2 + k_0^2)}
\nonumber \\
+ \frac{1}{2} \int \frac{d^3 p}{(2 \pi)^3} T \sum_{k_0} \int \frac{k^3}
{(2 \pi)^3} \left( \frac{g_w^2 \delta_{ab}}{2} i A_{\mu}^a(p)
i A_{\nu}^b(-p) \right)
\frac{i k_{\alpha} i (p+k)_{\beta} {\rm tr} ( \frac{1-\gamma^5}{2}
\gamma^{\alpha} \gamma^{\mu} \gamma^{\beta} \gamma^{\nu})}{(k^2 + k_0^2)
( (k+p)^2 + k_0^2)}
\end{eqnarray}
which corresponds to the Feynman diagrams illustrated in Fig. \ref{one}.
Higher order contributions can be gotten similarly by going to higher
powers in $H_I$.
The contribution of fermionic fluctuations to dimension two and four operators
in $SU(2) \times U(1)$ Higgs theory in $\overline{\rm MS}$
with realistic couplings
has recently been worked out by FKLRS \cite{FKLRS}. We independently
performed the calculation in $SU(2)$ Higgs theory; our results concur.
We have also extended the calculation to find all nonrenormalizeable operators
induced by fermions at dimension 6. Our results follow.
The dimension two operators induced by fermions change the masses of the
Higgs and $A_0$ fields. The induced terms are
\begin{equation}
\frac{N_c g_t^2 T^2}{12} \Phi^{\dag} \Phi
+ \frac{N_d g_w^2 T^2}{24} A_0^2 \, .
\end{equation}
The first term is largely responsible for the restoration of symmetry at
high temperature. The second is the familiar Debeye screening mass.
(Here and throughout $N_c = 3$ is the number of colors and $N_d = 12$ is
the number of left handed fermion doublets.)
At dimension four, fermions introduce the following corrections to couplings
and wave functions:
\begin{eqnarray}
N_c g_t^4 D4 \, (\Phi^{\dag} \Phi)^2
+ N_c g_t^2 D4 \, (D_i \Phi)^{\dag} D_i \Phi
+ \frac{N_c g_w^2 g_t^2}{4} (D4 - \frac{1}{8\pi^2}) A_0^2 \Phi^{\dag} \Phi
\nonumber \\
+ \frac{N_d g_w^2 D4}{12} F^a_{ij}F^a_{ij}
+ \frac{N_d g_w^2}{6}(D4 - \frac{1}{16\pi^2}) (D_i A_0)^a (D_i A_0)^a
- \frac{N_d g_w^4}{192 \pi^2} (A_0^2)^2 \, .
\label{dim4contrib}
\end{eqnarray}
The coefficients of $D4$ give the fermionic contributions to the one loop
beta functions and anomalous dimensions for the bosonic fields in $SU(2)$
Higgs theory, and agree with well known results. If the dimensional
reduction scheme is justified then these terms should give an accurate
approximation of the fluctuation determinant. To check this it is necessary
to go one order higher and see if the contributions of dimension 6
operators are small.
At dimension six, fermions contribute the following terms to the effective
potential,
\begin{equation}
-\frac{2 N_c g_t^6 D6}{3}(\Phi^{\dag} \Phi)^3
+ \frac{N_c g_t^2 g_w^4 D6}{16} (A_0^2)^2 \Phi^{\dag} \Phi \, ,
\end{equation}
the following Higgs derivative terms,
\begin{eqnarray}
- \frac{ 4 N_c g_t^4 D6}{3} (\Phi^{\dag} \Phi) (D_i \Phi)^{\dag} D_i \Phi
- \frac{ 2 N_c g_t^4 D6}{3} (\Phi^{\dag} D_i \Phi) (D_i \Phi)^{\dag} \Phi
\nonumber \\
- \frac{ N_c g_t^4 D6}{3} \partial_i(\Phi^{\dag} \Phi) \partial_i
(\Phi^{\dag} \Phi)
-\frac{N_c g_t^2 D6}{3} (D^2 \Phi)^{\dag} D^2 \Phi \, ,
\label{Higgsderivs}
\end{eqnarray}
the following mixed derivative terms,
\begin{equation}
\frac{N_c g_w^2 g_t^2 D6}{6} A_0^2 (D_i \Phi)^{\dag} D_i \Phi
- \frac{N_c g_w^2 g_t^2 D6}{12} (D_i A_0)^a (D_i A_0)^a \Phi^{\dag} \Phi \, ,
\end{equation}
the following $A_0$ derivative terms,
\begin{equation}
\frac{N_d g_w^4 D6}{24} \partial_i (A_0^2) \partial_i (A_0^2)
- \frac{N_d g_w^2 D6}{30} (D^2 A_0)^a (D^2 A_0)^a \, ,
\end{equation}
the following mixed terms,
\begin{equation}
- \frac{N_c g_t^2 g_w^2 D6}{8} F^a_{ij} F^a_{ij} \Phi^{\dag} \Phi
- \frac{N_d g_w^4 D6}{24} A_0^2 F^a_{ij} F^a_{ij}
+ \frac{N_d g_w^4 D6}{8} A_0^a F^a_{ij} A_0^b F^b_{ij} \, ,
\end{equation}
and the following gauge field terms,
\begin{equation}
-\frac{N_d g_w^3 D6}{180} f_{abc} F^a_{ij} F^b_{jk} F^c_{ki}
- \frac{ N_d g_w^2 D6}{15} (D_i F_{ij})^a (D_k F_{kj})^a \, .
\label{puregauge}
\end{equation}
Fermionic contributions to other independent gauge invariant dimension 6
operators (such as $(A_0^2)^3$) vanish.
Exactly the same terms arise whether the massive quark is top type or
bottom type. If both types are massive (in one generation) then each
occurrence of $g_t^n$ becomes $g_t^n + g_b^n$ in the above, and exactly one
mixed term appears at dimension 6,
\begin{equation}
- \frac{4 N_c g_t^2 g_b^2 D6}{3}
\Phi^{\dag} D_j i \tau_2 \Phi^* (D_j i \tau_2 \Phi^*)^{\dag} \Phi \, .
\end{equation}
Fermions also induce pure QCD operators and operators containing both weak and
strong fields; these are unimportant here and are listed in Appendix B.
We should comment that, except when they coincidentally vanish, the
fermionic contributions to dimension six operators tend to be much larger
than the contributions from the nonzero Matsubara frequencies of boson fields,
some of which have been worked out in \cite{Patkos}. This is partly because
the top quark is heavier than any of the bosonic degrees of freedom and
partly because its lowest nonzero Matsubara frequency is $\pi T$, rather
than $2\pi T$. Because of this the bosonic
equivalent of the coefficient $D6$ is smaller
by a factor of 7. Hence, we anticipate that if the expansion in high
dimension operators is well behaved for fermions then it will also be
well behaved for bosons.
We have also computed the fermionic contributions to masses, couplings, and
wave function renormalizations in a slight modification of the proper time
technique of DPSSG. The procedure is outlined in Appendix A. The
results turn out to be identical to those in $\overline{\rm MS}$ except
that $D4$ is modified to
\begin{equation}
D4_{\rm proper \: time} = \frac{ \gamma_E - 2 \ln \pi + \ln(\mu^2 / T^2)}
{16 \pi^2} \, ,
\label{D4propertime}
\end{equation}
where the definition of the scale $\mu$ is explained in Appendix A.
This expression relates the $\overline{\rm MS}$ and proper time renormalization
points. The appendix also presents the calculation of the vacuum effective
potential and the physical Higgs mass in the proper time scheme.
\section{The Sphaleron}
We want to apply these results to find the free energy of a nontrivial
field configuration which solves the classical equations of motion.
Klinkhammer and Manton have shown \cite{Klink} that the classical equations
of motion of the $SU(2)$ Higgs system can be solved by an {\it Ansatz}
of form
\begin{equation}
\Phi = \frac{\nu}{\sqrt{2}} \frac{h(r)}{r} \left[ \begin{array}{c} x+iy \\ z
\end{array} \right] \, , \qquad
A_i^j = \frac{2 f(r)}{r^2} r_k \epsilon_{ijk} \, .
\label{Sphaleron}
\end{equation}
The lower index on $A$ is the Lorentz index and the upper index is
the group index, and $h$ and $f$ are functions of $r$ alone which are
to be determined by minimizing the configuration's energy, subject to
the boundary conditions $f(0)=h(0)=0$ and $f(\infty)=h(\infty)=1$.
(Here and throughout $\nu$ is the Higgs VEV,
$\nu^2(T) = 2 \Phi^{\dag} \Phi (T) $.)
This solution is called the Sphaleron, and when $E_{Sph} >> T$, the
formalism of Langer tells us that most
baryon number violating events should occur because of phase space paths
which pass close to this configuration. Arnold and McLerran have applied
this idea to estimate the rate of baryon number violation in the broken
electroweak phase to be \cite{McLerran}
\begin{equation}
\frac{dN_B}{N_B dt} = -13 N_F T \left( \frac{\alpha_w}{4\pi} \right)^4
\frac{\omega_-}{2 m_W} \left( \frac{4 \pi \nu}{g_w T} \right)^7
{\cal N}_{\rm tr} {\cal NV}_{\rm rot} \kappa \exp(-E_{Sph}/T)
\label{eraserate}
\end{equation}
with $N_F=3$ the number of generations.
The prefactors arise from the zero and unstable modes of the Sphaleron,
and are evaluated in \cite{CarsonI,Carson}. At 1 loop $\kappa$ is
a product of fluctuation determinants around the configuration,
\begin{equation}
\kappa = \frac{ {\rm Det} H}{ {\rm Det} H_0} \left( \frac{ {\rm Det} K_0}
{ {\rm Det} K'} \right)^{\frac{1}{2}} \, .
\end{equation}
$H$ is Eq. (\ref{Hamiltonian}) in the Sphaleron background and $H_0$
is Eq. (\ref{Hamiltonian}) in the naive vacuum. $K_0$ is the bosonic
fluctuation determinant in vacuum, and $K'$ is the bosonic determinant in
the Sphaleron background, but with the zero and unstable modes removed.
We should comment that the division between $\ln \kappa$ and $-E_{sph}/T$ is
somewhat arbitrary. In particular it is renormalization point dependent.
We see this explicitly from our calculation of the contributions from
fermions to dimension 4 operators (Eq. (\ref{dim4contrib})), which are
part of $-\ln \kappa$, but which have
with coefficients $\propto D4$ which explicitly
depend on $\mu$. The calculation of $E_{Sph}$ correspondingly depends
on coupling constants $g_w$, $\lambda$ which depend on $\mu$.
To compute $E_{Sph} - T \ln \kappa$ we need to evaluate the tree level
Lagrangian terms $\Phi^{\dag} \Phi$, $(\Phi^{\dag} \Phi)^2$, and
$F^{a}_{ij} F^{a}_{ij}$, and the
operators found in Section II, in the Sphaleron
background, Eq. (\ref{Sphaleron}). First note that $A_0 = 0$, so all
terms including $A_0$ vanish. Effective potential terms give
\begin{eqnarray}
\Phi^{\dag} \Phi & = & \frac{\nu^2}{2} h^2,
\\
( \Phi^{\dag} \Phi )^2 & = & \frac{ \nu^4}{4} h^4,
\\
( \Phi^{\dag} \Phi)^3 & = & \frac{ \nu^6}{8} h^6 \ldots \: .
\end{eqnarray}
We will also need the term $ (\Phi^{\dag} \Phi)^{3/2} = \nu^3 h^3 /
(2 \sqrt{2})$ which arises from bosonic fluctuations, and we should add
a $\Phi^{\dag} \Phi$ independent constant to make the global minimum
of $V(\nu)$ zero, to subtract out the energy density in the absence of
a Sphaleron.
The dimension 4 derivative terms are \cite{Klink}
\begin{eqnarray}
F^a_{ij} F^a_{ij} & = & 16 \frac{f'^2}{r^2} + 32 \frac{f^2 (1-f)^2}{r^4} \, ,
\\
(D_i \Phi)^{\dag} D_i \Phi & = & \frac{\nu^2}{2} \left( 2 \frac{h^2 (1-f)^2}
{r^2} + h'^2 \right) \, .
\end{eqnarray}
We calculate that the dimension 6 derivative terms are
\begin{eqnarray}
f_{abc} F^a_{ij} F^b_{jk} F^c_{ki} & = & 96 \frac{f'^2 f (1-f)}{r^4} \, ,
\label{firstone} \\
(D_i F_{ij})^a (D_k F_{kj})^a & = & 8 \frac{f''^2}{r^2} - 32 \frac{f''
f(1-f)}{r^4} + 32 \frac{f^2 (1-f)^2}{r^6} \, , \\
\Phi^{\dag} \Phi F^a_{ij} F^a_{ij} & = & \frac{\nu^2 h^2}{2} \left(
16 \frac{ f'^2}{r^2} + 32 \frac{f^2 (1-f)^2}{r^4} \right) \, , \\
(D^2 \Phi)^{\dag} D^2 \Phi & = & \frac{\nu^2}{2} \left( h'' + 2 \frac{h'}{r}
-2 \frac{h(1-f)^2}{r^2} \right)^2 \, , \\
\Phi^{\dag} \Phi (D_i \Phi)^{\dag} D_i \Phi &=& \frac{\nu^4}{4} \left(
2 \frac{h^2 (1-f)^2}{r^2} + h'^2 \right) \, , \\
\partial_i (\Phi^{\dag} \Phi) \partial_i (\Phi^{\dag} \Phi) & = &
\nu^4 h^2 h'^2 \, , \\
\Phi^{\dag} D_i \Phi (D_i \Phi)^{\dag} \Phi & = & \frac{\nu^4}{4} \bigg(
\frac{h^4 (1-f)^2}{r^4} (x^2 + y^2) + h^2 h'^2 \bigg) \, .
\label{asphericalone}
\end{eqnarray}
Because $f \propto r^2$ and $h \propto r$ at small $r$,
all of these terms are nonsingular at $r = 0$.
The last term, and only the last term, is not spherically symmetric.
The departure from spherical symmetry arises because the Dirac equation
in the presence of
the Sphaleron is not spherically symmetric when only one flavor of quark
has a mass \cite{DPSSG}. DPSSG evaded this problem by giving both flavors
equal masses, which restores the symmetry. As we have seen, giving both
quark flavors masses introduces a new dimension 6 operator, whose
free energy density in the Sphaleron background is
\begin{equation}
\Phi^{\dag} D_i i \tau_2 \Phi^* (D_i i \tau_2 \Phi^*)^{\dag} \Phi =
\frac{\nu^4}{4} \frac{h^4(1-f)^2}{r^4} (r^2 + z^2) \, .
\label{extraterm}
\end{equation}
The coefficients of this term and (\ref{asphericalone}) are such that,
when $g_t^2 = g_b^2$, the $z^2$ combines with the $x^2 + y^2$ to restore
the spherical symmetry of the terms.
The approximation of DPSSG differs from the results with only the top quark
massive by the contribution of Eq.(\ref{extraterm}) and is accurate only
when this term is small. This is the case only when derivative dimension
6 operators give very small contributions, which is precisely the case where
dimensional reduction is accurate.
To compute the Sphaleron energy it is convenient to follow \cite{Klink} and
introduce a dimensionless radial distance $\xi = g_w \nu r$. The contribution
to the Sphaleron energy from the effective potential is then
\begin{equation}
\frac{4 \pi \nu}{g_w} \int_0^{\infty} d \xi \xi^2
\frac{V(h \nu)}{g_w^2 \nu^4} \, .
\label{Vcont}
\end{equation}
The contribution from kinetic energy terms, including the fermions'
contribution, is
\begin{equation}
\frac{4 \pi \nu}{g_w} \int_0^{\infty} \! \! d\xi \:
\left( 1 + \frac{ N_d g_w^2 D4}{3} \right) \left( 4 f'^2
+ 8 \frac{f^2(1-f)^2}{\xi^2} \right)
+ \left( 1 + N_c g_t^2 D4 \right) \left( h^2 (1-f)^2
+ \frac{ \xi^2 h'^2}{2} \right)
\label{derivcont}
\end{equation}
where derivatives are with respect to $\xi$.
The contribution from Eq. (\ref{puregauge}) is
\begin{eqnarray}
\frac{4 \pi \nu}{g_w} \frac{7 \zeta(3) g_w^4 \nu^2}{128 \pi^4 T^2}
\int_0^{\infty} d \xi \xi^2 \left(
- \frac{N_d}{180} \left[ 96 \frac{f'^2 f(1-f)}{\xi^4} \right] \right.
\qquad \qquad \qquad \qquad \qquad \nonumber \\
\left. - \frac{N_d}{15} \left[ 8\frac{f''^2}{\xi^2}
- 32 \frac{f'' f (1-f)}{\xi^4}
+ 32 \frac{f^2 (1-f)^2}{\xi^6} \right] \right) \, .
\label{puregaugecont}
\end{eqnarray}
We can get the other dimension 6 operators from Eqs.
(\ref{Higgsderivs} - \ref{puregauge})
and Eqs. (\ref{firstone} - \ref{asphericalone}) by always replacing
$D6$ with $7 \zeta(3) g_w^4 \nu^2/ 128 \pi^4 T^2$, removing $g_w$'s,
replacing $g_t$ with $g_t/g_w$ and $r$ with $\xi$, and integrating
over $ (4 \pi \nu / g_w) \int \xi^2 d\xi$. For $\nu \sim T$,
$D6$ is very small and dimension 6 operators have the parametric appearance
of two loop effects.
{}From the discussion after the introduction in \cite{McLerran} we see that
we want to find the configuration of form (\ref{Sphaleron}) with
minimum free energy, that is the configuration which maximizes
$(\Pi {\rm zero \; mode \: cont.}) \kappa \exp(-E_{Sph}/T)$. This is a
formidable task, since we cannot compute all of these terms analytically; in
particular the zero mode contributions and part of the zero Matsubara
frequency bosonic contribution to $\kappa$ are beyond our analytic abilities.
Fortunately, the Sphaleron is a saddlepoint configuration, and we should get
almost exactly the right free energy if we include the dominant effects
in the computation of the field configuration $f(\xi)$ and $h(\xi)$, and then
compute the other corrections holding $f$ and $h$ fixed. This is because,
as a saddlepoint, the free energy of the Sphaleron only changes
quadratically with small changes to $h$ and $f$.
To illustrate this point consider the Sphaleron configuration
computed from a tree level
Lagrangian $ F^{a}_{ij} F^{a}_{ij}/4$$
+ \lambda (\Phi^{\dag} \Phi - \nu^2 /2 )^2$ and
suppose that we are only interested in the correction fermions induce in the
gauge fields, $N_d g_w^2 D4 F^{a}_{ij} F^{a}_{ij}/12$.
We will choose $\mu = m_W(T=0)$
in the proper time scheme and $T = 100$ GeV, so $D4 \simeq - 0.0135$.
Solving for the Sphaleron using the tree action we find
$E_{Sph} = 35.0975 \nu$,
and estimating the fermionic correction as
\begin{equation}
\frac{4 \pi \nu}{g_w} \int_{0}^{\infty} d \xi \left( \frac{N_d g_w^2 D4}{3}
\right) \left( 4 f'^2 + 8 \frac{ f^2 (1-f)^2}{\xi^2} \right)
\simeq -0.4433 \nu
\end{equation}
we get a total free energy of $34.6542 \nu$. When we solve for the Sphaleron
configuration, including the fermionic contribution as well as the tree
terms, we find the Sphaleron energy, which now includes the fermionic
correction, is $34.6520 \nu$. The correction from this fermionic
contribution is about $1\%$ of $E_{Sph}$, and the error in estimating it
at fixed $f$ and $h$ is about $0.01 \%$, which is quadratic, as expected.
Obviously, though, this will not do when a correction is actually
substantial and the modification of the free energy is comparable to
$E_{Sph}$. This is potentially the case for the fermionic correction to
the $(\Phi^{\dag} \Phi)^2$ term in the effective potential. For the
above parameters, the coefficient of the correction is $N_c g_t^4 D4
= -0.040$, which is to be compared with $\lambda(\mu) = 0.050$.
Because of an unfortunate choice for $\mu$, the ``correction'' for
top quarks is almost as large as the tree level term itself, and we
certainly cannot trust a result in which $f$ and $h$ are computed with
the tree term and used to compute the quark fluctuation determinant.
The only consistent,
renormalization point independent thing to do is to include those
fermionic contributions which correct operators appearing in the tree
level action in the calculation of the Sphaleron configuration,
that is to use the $\mu$ independent quantity $\lambda(\mu) + N_c g_t^4 D4$
in the calculation of $f$, $h$, and $E_{Sph}$. This is particularly
important for the scalar self-coupling, because as mentioned earlier it
is unprotected from large radiative corrections and in fact the top
quark contribution here is substantial.
To work in terms of physical quantities, we first find a relation
between the couplings at the scale $\mu$ and vacuum, physical masses. That
is, we find $\lambda(\mu)$ in terms of $m_H(T=0)$. We perform this
calculation, including all one loop, fermionic contributions, in
Appendix A. We ignore the bosonic corrections because we are most
interested in understanding the fermionic radiative corrections in this
paper, and because the bosonic corrections are smaller by a factor
of $9 m_W^4/12 m_t^4 \simeq 1/30$. From the Appendix A results we find
$\lambda ( \mu ) = (m_H^2/2 \nu_o^2)
- (N_c g_t^4/16 \pi^2) (\ln (\mu^2/m_t^2) - a)$,
where $a=0$ in $\overline{\rm MS}$ and $a = \gamma_E$ in proper time
regulation.
The quantity we should use in calculating the Sphaleron
configuration at a temperature $T$ is
\begin{equation}
\lambda_T \equiv \lambda(\mu) + N_c g_t^4 D4 \, ,
\label{shoulduse}
\end{equation}
which has no $\mu$
or renormalization scheme dependence, as we can note by inspecting
Eq. (\ref{D4def}) or Eq. (\ref{D4propertime})\footnote{The alert
reader may notice that $\lambda(\mu)$ also contains a term
$(m_H^2/2\nu_o^2)( N_c g_t^2/16 \pi^2)( \ln (\mu^2 / m_t^2) - a)$, which
apparently spoils the cancellation of the $\mu$ dependence discussed here.
This is true, and it gets another correction because $\nu_o$ is $\mu$
dependent;
but the mass squared parameter $m_0^2$ of the effective
potential contains a similar $\mu$ dependence, so the effect of the correction
is just to shift the location of the minimum of $V$ by a proportional
amount, without changing its height.
The Higgs field wave function has the same dependence, so this has no
influence on the Sphaleron energy.}.
A particularly convenient choice for
$\mu$ is
\begin{equation}
\mu = \exp(- \gamma_E) \pi T, \; \overline{\rm MS} \quad {\rm or}
\quad
\mu = \exp(- \gamma_E / 2) \pi T, \; {\rm proper \; time}
\label{muchoice}
\end{equation}
because in this case $\lambda_T = \lambda(\mu)$ and the Higgs field and
gauge field wave functions take their tree values; but there is no need
to choose this scale, as long as we compute the Sphaleron configuration
and energy using couplings and wave functions corrected by the fermion
contributions as in equations (\ref{derivcont},\ref{shoulduse}).
This is precisely the
prescription of the dimensional reduction program. We can test its
reliability by seeing how large the remaining corrections, those coming
from dimension 6 operators, are.
Of course, it is also necessary to
include corrections from zero Matsubara frequency bosons, which are
expected to be considerable and cannot be evaluated with a derivative
expansion, as we have already discussed. The largest part of this
correction comes from the cubic effective potential terms. The residual
correction when this is completely removed is $\ln \kappa \simeq 1.5$
\cite{Junker,Junker2} which is small enough that we can trust the
computation in terms of fixed $f$ and $h$ as discussed above; but
the effective potential contribution is large (at $T \simeq T_c$ it
changes the very nature of the phase transition and should be considered
an order 1 correction to $E_{Sph}$) and should be included in the
evaluation of the Sphaleron configuration, as advocated in \cite{Shap}.
But before we can compute
the Sphaleron energy including these corrections we must discuss the
nature of the phase transition to find out at what temperature the
baryon erasure begins.
\section{Phase transition, Sphaleron energy}
Let us briefly review the electroweak phase
transition. At one loop the zero Matsubara frequency bosonic excitations
generate negative cubic terms in the effective potential, which becomes
($\Theta_W = 0$)
\cite{Arnold}
\begin{eqnarray}
V( \nu ) = - \frac{g_* \pi^2 T^4}{90} +
\left( - \frac{m_0^2}{2} + \frac{(4g_t^2+3 g_w^2 + 8 \lambda)T^2}{32}
\right) \nu^2
- \frac{g^3 T}{16 \pi} \nu^3
- \frac{ g_w^3 T}{4 \pi}
\left( \frac{11 T^2}{6} + \frac{\nu^2}{4} \right)^{\frac{3}{2}}
\nonumber \\
- \frac{3 m_2^3 + m_1^3}{12 \pi} + \lambda_T \frac{\nu^4}{4}
+{\rm dimension \; six} \, , \quad
\end{eqnarray}
where $m_1^2 = \lambda_T \nu^2/2 + V''(\nu = 0)$,
$m_1^2 = \lambda_T \nu^2 + m_2^2$, $g_* = 106.75$ is the number
of radiative degrees of freedom, and ``dimension six'' means the higher terms
found in Eq. (\ref{hiTexpan}).
At high temperature $V$ has only one minimum
at $\nu = 0$, but as temperature drops the negative cubic terms generate
a second ``asymmetric'' minimum. At some temperature the second minimum
becomes more thermodynamically favorable, and bubbles of the asymmetric
phase begin to nucleate and grow shortly thereafter. The temperature at
which the nucleations become common can be computed by standard techniques
\cite{Dine}. The expanding bubbles liberate latent heat, so that the
temperature after the transition is somewhat higher than the nucleation
temperature. The temperature the plasma reheats to due to this latent heat
is determined by the condition that the broken phase energy density
$E = V - T \partial V / \partial T$
must equal the symmetric phase energy density at the temperature where
the nucleations occurred.
It is at this reheat temperature, immediately after the phase
transition, that quasi-equilibrium erasure of any baryon number excess
generated at the phase transition begins. It continues for all times
thereafter, but as Eq. (\ref{eraserate}) shows,
the rate depends strongly on the
Sphaleron energy, which changes rapidly with $T$, as $\nu$ moves towards
its zero temperature value. Almost all the baryon erasure takes place within
a fraction of a Hubble time, so the Sphaleron
rate is only relevant at temperatures quite close to the reheat temperature.
We have computed the reheat temperature and the Sphaleron energy at the
reheat temperature for a number of physical Higgs masses, using the 1
loop relations between Higgs mass and $\lambda$ presented in Appendix A and
including the negative cubic effective potential terms from zero Matsubara
frequency bosonic modes.
(We always use $g_w = 0.65$ and take the Weinberg angle to be zero. In fact,
to account for the quark contribution to the gauge
field wave function, the value of $g_w$ we should use is
$g_w(\mu) - N_d g_w^3 D4/6$, which is $T$ dependent. The $T$ dependence
is very mild, behaving as $g_w * (N_d g_w^2 / 48 \pi^2 \simeq 0.01) \ln T$.
Since this dependence is so weak, and
since we are not including the influence of bosons' nonzero Matsubara modes,
which will contribute an opposite and slightly larger temperature dependence
\cite{FKRS}, we will not worry about it. However we will take full
consideration of the correction to $\lambda$, which as discussed is not at
all weak.)
The results, together with the contributions from
dimension 6 operators, are presented in table 1. The $\nu^6$ contribution
to the effective potential was computed by inclusion in the calculation of
the Sphaleron configuration, and the others were performed perturbatively.
It is clear from the table that, as expected,
dimension 6 operators make only a tiny contribution to the Sphaleron energy.
Hence we can conclude that near the phase transition temperature the
expansion is very well behaved and the use of the dimensionally reduced
theory is well justified. The largest dimension 6 correction comes from
the effective potential term, owing to the high power of $g_t$; this and
only this term may not be completely negligible, increasing
the Sphaleron energy and the strength of the phase transition
by a few percent for light Higgs.
Generally the terms become progressively less important as they contain more
derivatives. In particular the very small contribution from
$(D^2 \Phi)^{\dag}) D^2 \Phi$ gives confidence that the expansion of
Eq. (\ref{derivexpan}) in derivatives is justified.
We have not continued the table down below 30 GeV partly because this
range is experimentally excluded, partly because questions of vacuum
stability become increasingly hard to avoid in this range, and partly
because the results barely differ from those at $m_H = 30$ GeV.
This is because the one
loop relations for the Higgs mass give nonzero $\lambda_T$ even as the
physical Higgs mass $m_H \rightarrow 0$.
The next step is to use these results to determine the baryon number
depletion. From Eq.(\ref{eraserate}) we find that baryon number is depleted
by a factor of
\begin{eqnarray}
\exp \left(
\int_{t_{\rm reheat}} 13N_F T \left( \frac{\alpha_w}{4 \pi} \right)^4
\frac{\omega_-}{2 m_W} {\cal N}_{tr} {\cal NV}_{rot}
\left( \frac{4 \pi \nu}{g_w T} \right)^7 \kappa \exp \frac{-E_{Sph}}{T} dt
\right)
\label{erasure}
\end{eqnarray}
The elapsed time is related to the change in temperature by $dt = dT/(HT)$,
with $H$ the Hubble constant, which is approximately $H = T(8 \pi^3 g_*/90)
^{1/2} (T/m_{pl})$. To perform the integral
we must numerically repeat the evaluation of the
Sphaleron energy at many values of $T$ close to the reheat temperature.
Most of the coefficients in Eq. (\ref{erasure}) are given in
\cite{CarsonI}. We use the values found there for $\omega_-$ and
the ${\cal N}$, even though they were computed for slightly different $f$
and $h$, because the product of
these factors turns out to be very insensitive
to changes in the configuration \cite{CarsonI}.
The other factor we require is the difference between the
Sphaleron energy which we have calculated and the (1 loop) value of
$E_{Sph} - T \ln \kappa$. This is due to derivative corrections from
zero Matsubara frequencies and can be approximated from the results
of Baacke and Junker \cite{Junker}, who find that, when the full tadpole
is removed (ie. all effective potential contributions are subtracted out),
$\ln \kappa$ is about 1.5, independent of $\nu$ and quite weakly dependent
on $\lambda_T$. (Again, this quantity was computed for different $f$, $h$ in
that paper, but again it proved to be quite insensitive to configuration,
so we will use this value. This introduces an uncertainty in our final value
for the erasure rate of perhaps $\pm 1$ in the exponential.)
We can then perform the integral in
Eq. (\ref{erasure}) numerically. We find that almost all of the
erasure occurs in a range of $T$ less than $0.5 \% $ from the reheat
temperature, essentially because $\nu$ changes very rapidly with temperature
immediately after the transition. This narrows the available time for the
Sphaleron erasure, but even for
$m_H =$30 GeV we find that the baryon number is depleted by a factor of
about $\exp(11.3)$. If the conjecture about the effect of Landau damping
on the negative frequency mode in \cite{McLerran} is correct then the
suppression is smaller by about 2 in the exponent.
The results for several Higgs masses are presented
in the table; in all cases the erasure is very considerable. Since the
most optimistic estimates of baryogenesis in the minimal standard model
can barely account for the current abundance \cite{Farrar}, this
apparently rules out electroweak physics as the source of baryogenesis
in the minimal standard model.
\section{Conclusion}
It appears that a perturbative treatment of fermions is very well justified
near the phase transition temperature,
and that dimensional reduction should be accurate, though the correction
{}from the dimension 6 contribution to the effective potential may not be
completely negligible.
How should we understand the results of DPSSG in
light of this conclusion?
First recall what we have done here. We find
that at a general $\mu$ the fermions will induce nonzero corrections to
couplings and wave functions, and in particular the correction to $\lambda$
is potentially large. Following the idea of the dimensional reduction
program we combine these contributions with the tree couplings, resulting
in renormalization point independent couplings $\lambda_T$, which are used
to compute the Sphaleron energy. The Sphaleron energy then already contains
that part of the fermion fluctuation determinant which is understood as
coupling and wave function corrections; the residual correction, which
comes from dimension 6 (and higher) operators, is explicitly found to be
very small. However, had we used the tree couplings at some scale $\mu$,
we would then expect fermions to give a ($\mu$ dependent) nonzero
correction due to the difference $\lambda_T - \lambda(\mu)$. Only at
one particular, convenient renormalization point (Eq. (\ref{muchoice}))
would this contribution vanish.
Diakonov et. al. use a fixed $\mu$ (in \cite{DPSSG} they use $\mu=m_W$
in the proper time scheme,
and in \cite{DPSSG2} they use $\mu \simeq m_t \exp(\gamma_E/2)$) and
compute the Sphaleron configuration from the tree Lagrangian (corrected
however by the thermal contributions to the $\Phi^{\dag} \Phi$ term). We
then expect from our work that they should find a correction arising
from the difference $\lambda_T - \lambda(\mu)$ equal to
\begin{equation}
\frac{4 \pi \nu}{g_w} \frac{\lambda_T - \lambda(\mu)}{g_w^2}
\int_{0}^{\infty} d\xi \xi^2 \frac{h^4 - 1}{4} \, ,
\end{equation}
which will depend on the top mass as $g_t^4$, exactly as they find. This is
not a contradiction of the dimensional reduction scheme, which predicts
that, because of their use of $\lambda(\mu)$ with their choice of $\mu$,
they should find such a term. We should note that, when this term is
large (as it is for the choice of $\mu$ made in \cite{DPSSG}), one
should not trust the functions $h$, $f$ computed from the tree
Lagrangian but should include this radiative correction in their computation,
as we have done here. In other words, the numerical work of
Diakonov et. al. is probably accurate, but because of the way they
have done the problem they will not necessarily produce the free energy
of the configuration which actually limits the baryon erasure rate.
We should also note that in \cite{DPSSG} Diakonov et. al.
use a tree, rather than 1 loop,
relation between $m_H$ and $\lambda(\mu)$, and a tree, rather than 1 loop,
value of $\lambda$ in calculating
the phase transition temperature. (In \cite{DPSSG2} the
relation between $m_H$ and $\lambda(\mu)$ is computed at one loop, but the
transition temperature is still found using $\lambda(\mu)$ rather than
$\lambda_T$.) This is inconsistent with a one loop analysis of the
Sphaleron rate and may explain the difference in our
results for the dissipation of baryon number.
What is the overall effect of fermions? For small Higgs mass,
$\lambda_T > \lambda_{\rm tree}$,
and as $\nu(T_{\rm reheat})$, and hence $E_{Sph}$, fall with increasing
$\lambda_T$, we find more baryon number dissipation than we would if we
ignored fermions altogether and used $\lambda_{\rm tree}$. This is the
reason that, even for very small Higgs mass, we still find substantial
baryon number erasure. We should emphasize once more that to find this
result it was important to apply one loop corrections systematically, in
the effective potential, the phase transition temperature, and the Sphaleron
energy, but that the effect is basically perturbative, and the high
temperature expansion accounts for it successfully.
Do the results of the last section preclude electroweak baryogenesis?
They make it unlikely that baryogenesis can be viable in the minimal
standard model. However we should note that we have only used the 1 loop
effective potential, and while extending our results to the two loop
potential, which is known, gives essentially the same conclusions, it is
not clear that the perturbative treatment of the effective potential is
reliable; the phase transition may be stronger than perturbation theory
suggests. Also, we have said nothing about extensions to the standard model.
For instance, in the two doublet model, the phase transition can be much
stronger without contradicting experimental bounds on the Higgs mass
\cite{someone}, and the Sphaleron bound only narrows the parameter space.
\centerline{Acknowledgements}
I am very grateful to Misha Shaposhnikov for useful conversations,
correspondence, and encouragement, and to Mikko Laine, for comparing
unpublished calculations of the fermionic contributions to dimension 4
operators with me. I also acknowledge funding from the NSF.
\section{Appendix A: Proper Time Calculations}
In this appendix we will show how to compute the finite temperature
fluctuation determinant using a proper time regulation analogous to that of
DPSSG. The basic idea is to use the relationship
\begin{equation}
\ln \frac{ {\rm Det} K}{ {\rm Det} K_0} = \lim_{\epsilon \rightarrow0}
- { \rm Tr} \int_{\epsilon}^{\infty} \frac{dt}{t}
(e^{-Kt} - e^{-K_0t}) \, ,
\end{equation}
which holds when $K$ is a quadratic, positive definite operator and $K_0$
is its free, massless approximation. In our case the operator $\gamma^\mu
\partial_\mu - m$, or Eq. (\ref{Hamiltonian}), is not quadratic or positive
definite. Fortunately it has the same spectrum as the operator
$-\gamma^\mu \partial_\mu - m$, so we can write the determinant we actually
want in terms of a quadratic, positive definite operator,
\begin{eqnarray}
-\ln {\rm Det}( \gamma^\mu \partial_\mu - m)
& = & -\frac{1}{2} \left(
\ln {\rm Det}( \gamma^\mu \partial_\mu - m)
+ \ln {\rm Det} ( -\gamma^\mu \partial_\mu - m) \right)
\nonumber \\
& = & - \frac{1}{2} {\rm Tr} \ln ( -\gamma^\mu \gamma^\nu \partial_\mu
\partial_\nu - \gamma^\mu (\partial_\mu m) + m^2) \, ,
\end{eqnarray}
which acts on a Euclidean space where the time direction is cylcic with
antiperiodic boundary conditions and period $T$. It therefore already
includes all thermal effects, which do not have to be added by hand,
as in the treatment of DPSSG. However, we will have to be careful
when we regulate to see that we are making $T$ independent subtractions.
Following \cite{DPY} we perform the trace by inserting a complete set
of plane-wave states,
\begin{equation}
{\rm Tr} e^{-Kt} = {\rm tr} \int d^4 x T \sum_{p_0} \int \frac{d^3
p}{(2\pi)^3} e^{-i p \cdot x} e^{-Kt} e^{i p \cdot x} \, ,
\end{equation}
where tr is over Dirac indicies. The factor $\exp(i p \cdot
x)$ can be brought through the operator to cancel $\exp(-ip \cdot x)$,
but in doing this all derivative operators in $K$ are shifted,
$\partial_\mu \rightarrow \partial_\mu + i p_\mu$. The zero
temperature limit is found by replacing the sum on $p_0$ with the
integral $\int dp_0/(2 \pi)$.
We illustrate the renormalization procedure by computing the effective
potential in this regulation. In this case $K = - \partial^2 + m^2$
because the mass is space independent. After the shift, the
derivatives have nothing on which to act and do not contribute; we can
drop them. The problem is then to compute
\begin{equation}
2 \int_{\epsilon}^{\infty} \frac{dt}{t} \int d^3x
\sum_{p_0} \int \frac{d^3 p}{(2 \pi)^3}
e^{- p^2 t} e^{-p_0^2 t} (e^{-m^2 t} - 1) \, ,
\end{equation}
where we have performed the Dirac trace and removed the trivial integral
over $dx_0$ so that the
result will be the free energy density.
Next we expand $\exp(-m^2 t)$ in powers of $t$. The first term is
cancelled by the $-1$. The second gives
\begin{equation}
-2\int m^2 d^3 x \int_{\epsilon}^{\infty} dt T \sum_{p_0} e^{- p_0^2 t} \int
\frac{d^3 p}{(2 \pi)^3} e^{-p^2 t} \, .
\end{equation}
This expression is small $t$ divergent, cut off by $\epsilon$. To
render the theory cutoff independent we should add and subtract a
temperature independent expression with the same small $t$ behavior
and absorb the one we added with a counterterm in the tree level mass.
The correct expression to subtract is
\begin{equation}
-2 \int m^2 d^3 x \int_{\epsilon}^{\infty} dt \int \frac{dp_0}{2 \pi}
e^{-p_o^2 t} \int \frac{d^3 p}{(2\pi)^3} e^{-p^2 t} \, .
\end{equation}
There is now no obstacle to performing the integral over $t$, or to
setting $\epsilon$ equal to zero. The result is
\begin{equation}
-2\int m^2 d^3 x \int \frac{d^3 p}{(2 \pi)^3} \left( T \sum_{p_0}
- \int \frac{dp_0}{2 \pi} \right) \frac{1}{p^2 + p_0^2} \, .
\end{equation}
Both the sum and the integral over $p_0$ are straightforward, giving
\begin{equation}
\sum_{p_0} \frac{1}{p_0^2 + p^2} = \frac{ {\rm tanh}
\frac{p} {2T}}{2p} \, , \quad \int \frac{dp_0}{2\pi} \frac{1}{p_0^2 + p^2}
= \frac{1}{2p} ', ;
\end{equation}
combining them and performing the integral over angles, we get
\begin{equation}
-2 \int d^3x m^2 \frac{1}{2 \pi^2} \int \frac{-1}{\exp(p/T) + 1} p dp
= \int d^3 x \frac{m^2 T^2}{12} \, ,
\end{equation}
which is the well known expression for the thermal contribution to the
Higgs mass squared.
In their discussion of the renormalization of the theory DPSSG
advocate cutting off the proper time integral of the counterterm at
some finite upper bound, call it $\mu^{-2}$ (in their case, $m_W^{-2}$).
Doing so changes the result of the above calculation to
\begin{equation}
\frac{m^2 T^2}{12} - \frac{\mu^2 m^2}{8 \pi^2} \, ,
\end{equation}
which means there is a discrepancy between the tree level mass squared
parameter and the renormalized, vacuum mass squared parameter. There
is nothing wrong with this in principle as long as we remember it is
there; we should use the renormalized mass squared when performing
calculations such as the Sphaleron configuration and remember that we
have already included part of the fermion contribution by doing so; we
will need to subtract it off, along with the thermal mass squared. It
is much easier and more straightforward, however, to follow our
procedure and absorb the entire vacuum correction in a counterterm.
Continuing to expand $\exp(-m^2 t)$, the next term is
\begin{equation}
+\int m^4 d^3x \int_{\epsilon}^{\infty} t dt \sum_{p_0}
e^{-p_0^2 t} \int \frac{d^3 p}{(2\pi)^3} e^{-p^2 t} \, ,
\end{equation}
which is logarithmically divergent at small $t$. Because the lowest
Matsubara frequency is nonzero, it is cut off exponentially at large
$t$ and there are no infrared problems in its evaluation. Again we
have an available Lagrangian parameter in which to absorb the
divergence, and we should add and subtract a $T$ independent
expression with the same ultraviolet behavior, and absorb the one with
the same sign into the Higgs self-coupling parameter. The
corresponding vacuum integral is both ultraviolet and infrared
divergent, and to prevent infrared divergences we
must introduce a renormalization scale into the problem. Following
DPSSG, we subtract
\begin{equation}
\int m^4d^3 x \int_{\epsilon}^{\mu^{-2}} t dt \int \frac{dp_0}{2 \pi}
e^{-p_0^2 t} \int \frac{d^3 p}{ (2\pi)^3} e^{-p^2 t} \, ,
\end{equation}
giving
\begin{equation}
\int m^4 d^3x \left( \int_{\epsilon}^{\infty} dt T \sum_{p_0} -
\int_{\epsilon}^{\mu^{-2}} dt \int \frac{dp_0}{2\pi} \right)
t e^{-p_o^2t} \int\frac{ d^3 p}{ (2\pi)^3} e^{-p^2 t} \, .
\end{equation}
The integral over $d^3p$ gives $t^{-3/2}/( 8 \pi^{3/2})$. The
counterterm can then be evaluated directly and gives $+ \ln( \mu^2
\epsilon ) / (16 \pi^2)$. The remaining integral,
\begin{equation}
\frac{T}{8 \pi^{ \frac{3}{2}}}\sum_{p_0} \int_{\epsilon}^{\infty}
t^{- \frac{1}{2}} dt e^{-p_0^2 t} \, ,
\label{toughint}
\end{equation}
is more delicate. For small terms in the sum the integral over $t$ is
approximately $\sqrt{\pi} / p_0$, so the early terms in the sum are
\begin{equation}
\frac{T}{4 \pi} \sum_{l=1,3...} \frac{1}{l \pi T} \, .
\end{equation}
The difference between this sum and a corresponding integral is
concentrated in the first few terms. The sum is approximately
\begin{equation}
\frac{1}{4 \pi} \left[ \left( \int_{\pi T} \frac{dp_0}{2\pi}
\frac{1}{p_0} \right) + \frac{ \gamma_E + \ln 2}{2 \pi} \right] \, ,
\end{equation}
so smoothing over the sum in Eq. (\ref{toughint}) introduces a
correction of $(\gamma_E + \ln 2)/(8 \pi^2)$, giving
\begin{equation}
\frac{\gamma_E + \ln 2}{8 \pi^2} + \frac{1}{4 \pi^{\frac{3}{2}}} \int
_{\pi T}^{\infty} \frac{dp_0}{2 \pi p_0} \int_{\epsilon}^{\infty} \frac
{dt} {t^{\frac{1}{2}}} e^{-p_0^2 t} \, .
\end{equation}
The integral over $p_0$ can be performed by parts, giving
\begin{equation}
\frac{\gamma_E + \ln 2}{8 \pi^2} + \frac{1}{4 \pi^{ \frac{3}{2}}}
\frac{-1} {2 \sqrt{\pi}} \left( \ln \pi T + \frac{1}{2} \ln \epsilon -
\frac{1}{2} \psi(1/2) \right) \, ,
\end{equation}
with $\psi$ the digamma function, $\psi(1/2) = -\gamma_E - 2 \ln
2$. The result is then
\begin{equation}
\int m^4 d^3 x \frac{ \gamma_E - 2 \ln \pi + \ln (\mu^2/T^2)}{16\pi^2} \, .
\end{equation}
The fraction in the integral gives the proper time renormalization value
of $D4$.
The terms higher order in $m^2$ are actually easier; the term at order
$m^{2n}$ is given by the integral
\begin{equation}
2 \int_{0}^{\infty} \frac{dt}{t} \int d^3 x \frac{(-1)^n}{n!} t^n m^{2n}
\sum_{p_0} \int \frac{d^3 p}{(2 \pi)^3} e^{-p^2 t} e^{-p_0^2 t}\, .
\end{equation}
Performing the integral over $t$, we get
\begin{equation}
\frac{2 (-1)^n}{n} \int d^3 x m^{2n} \sum_{p_0}
\int_{0}^{\infty} \frac{p^2 dp}{2 \pi^2} \frac{1}{(p^2 + p_0^2)^n} \, .
\end{equation}
The result per unit volume is
\begin{equation}
\frac{ (-1)^n m^{2n}}{n \pi^2} \left( \sum_{p_0} p_0^{3-2n} \right)
\int_{0}^{\infty} \frac{y^2 dy}{(y^2 + 1)^n} \, ,
\end{equation}
from which Eq. (\ref{hiTexpan}) follows immediately by performing the sum
and the integral. Note that the calculation was completely large $t$ finite
because the $\exp(-p_0^2 t)$ term always decays exponentially fast, since
$p_0^2$ is always at least $(\pi T)^2$. This is to be contrasted with
the zero temperature \cite{DPY} and zero Matsubara frequency bosonic
\cite{CarsonI} cases and explains why the expansion in operator dimension
is possible at finite temperatures for fermions, while it is known to
have trouble at zero temperature and for the zero Matsubara frequency bosonic
modes.
The same basic techniques can be used for the case with gauge fields
and spatial variation.
A complication which arises when computing other dimension 4 operators
is that, in contributions involving $A_0$, powers of $p_0$
appear in addition to $\exp(-p_0^2 t)$. The result is that the
argument of the digamma function shifts by 1 per power of $p_0^2$. This
is why the dimension 4 terms containing $A_0$ are not simply multiples of $D4$.
Next we will compute the fermionic contribution to the vacuum
effective potential in this regulation. From the above,
the contribution from each color is
\begin{equation}
2 \left[ \left( \int_{\epsilon}^{\infty} \frac{dt}{t}
\int \frac{d^4 p}{(2 \pi)^4} e^{-p^2 t} (e^{-m^2 t} - 1 + m^2 t) \right)
- \left( \int_{\epsilon}^{\mu^{-2}} \frac{dt}{t} \int \frac{d^4p}{(2\pi)^4}
e^{-p^2 t} \frac{m^4}{2} t^2 \right) \right] \, .
\end{equation}
The $-1$ is the vacuum energy subtraction, the $m^2 t$ is the mass squared
counterterm, and the last expression is the self-coupling counterterm.
The integrals over $p$ may be performed immediately, giving
\begin{equation}
\frac{1}{8\pi^2} \left(\int_{\epsilon}^{\infty} \frac{dt}{t^3}
(e^{-m^2 t} - 1 + m^2 t) \right)
- \frac{1}{8 \pi^2} \frac{m^4}{2} \int_{\epsilon}^{\mu^{-2}} \frac{dt}{t} \, .
\end{equation}
After integrating the first expression by parts three times, we get
\begin{equation}
\frac{m^4}{16 \pi^2} \left( \frac{3}{2} + \ln \frac{\mu^2}{m^2} - \gamma_E
\right) \, .
\label{Veffpropertime}
\end{equation}
Recalling that $m^2 = g_t^2 \Phi^{\dag} \Phi$, summing on colors, and adding
this expression to the tree level effective potential
\begin{equation}
-m_0^2 \Phi^{\dag} \Phi + \lambda (\Phi^{\dag} \Phi)^2
\end{equation}
we find that the curvature of the effective potential (the second derivative
with respect to $\nu$) at the minimum ($\nu = \nu_o$) is
\begin{equation}
V'' = 2 \nu_o^2 \left( \lambda + \frac{N_c g_t^4}{16 \pi^2}
( \ln \frac{\mu^2}{m_t^2} - \gamma_E ) \right) \, ,
\label{Vdubprime}
\end{equation}
and the effective potential parameter $m_0^2$ is
\begin{equation}
m_0^2 = \frac{V''}{2} + \frac{N_c g_t^4 \nu^2}{16 \pi^2} \, .
\end{equation}
The analogous $\overline{\rm MS}$ result is the same but with the
$- \gamma_E$ removed from Eq. (\ref{Vdubprime}).
$V''$ is not the physical Higgs mass squared.
The physical Higgs mass is the ratio
of potential to kinetic energy of a long wavelength fluctuation about the
minimum, times the wave number of the fluctuation; there is a wave function
correction which requires computing the self-energy at the pole mass.
We have been unable to do the calculation in proper time regulation
(see however \cite{DPSSG2}), but we have performed it in $\overline{\rm MS}$
using an integral from \cite{Bohm}. The result is that
\begin{equation}
m_H^2 = V'' \left( 1 - \frac{N_c g_t^2}{16 \pi^2}
\bigg( \ln \frac{ \mu^2}{m_t^2} + 2
- \frac{2 \sqrt{4 m_t^2 - m_H^2}}{m_H} {\rm arctan} \frac{m_H}
{\sqrt{4 m_t^2 - m_H^2}} \; \bigg) \right) \, .
\label{mHpropertime}
\end{equation}
The proper time value is presumably the same but with a $-\gamma_E$ inserted
after the log. For $4m_t^2 >> m_H^2$ the expression following the log is
about $2 m_H^2/3(4m_t^2 - m_H^2) \simeq 0$ and we have permitted ourselves
to neglect it when relating the quartic coupling to the physical Higgs mass.
\section{Appendix B: Other high dimension operators}
We list here those dimension 4 and 6 operators induced by fermions
which contain both weak $SU(2)$ and strong $SU(3)$ fields.
These are probably of no phenomenological consequence and
almost certainly do not influence the strength of the phase
transition. We include them only for completeness.
We write the time component of the gluon field as $A_{0g}$ and the
field strength tensor as $G_{ij}$. Denoting the number of families
as $N_F = 3$, the dimension four, mixed terms are
\begin{equation}
\frac{-g_t^2 g_s^2}{8 \pi^2} A_{0g}^2 \Phi^{\dag} \Phi
- \frac{g_w^2 g_s^2 N_F}{16 \pi^2} A_0^2 A_{0g}^2 \, .
\end{equation}
The dimension six mixed terms are
\begin{eqnarray}
2 g_t^2 g_s^2 D6 A_{0g}^2 (D_i \Phi)^{\dag} D_i \Phi
+ g_t^4 g_s^2 D6 A_{0g}^2 (\Phi^{\dag} \Phi)^2
+ \frac{g_s^2 g_w^2 g_t^2 D6}{8} A_{0g}^2 A_0^2 \Phi^{\dag} \Phi
\nonumber \\
+ \frac{g_s^2 g_w^2 N_F D6}{3} A_{0g}^2 (D_i A_0)^a (D_i A_0)^a
+ \frac{g_s^2 g_w^2 N_F D6}{6} A_{0g}^2 F^a_{ij} F^a_{ij}
\nonumber \\
+ \frac{g_s^2 g_w^2 N_F D6}{3} (D_i A_{0g})^a(D_i A_{0g})^a A_0^2
+ \frac{2 g_s^2 g_t^2 D6}{3} \partial_i (A_{0g}^2)
\partial_i (\Phi^{\dag} \Phi)
\nonumber \\
+ \frac{g_s^2 g_w^2 N_F D6}{3} \partial_i(A_{0g}^2) \partial_i (A_0^2)
- \frac{g_s^2 g_t^2 D6}{3} G^a_{ij} G^a_{ij} \Phi^{\dag} \Phi
+ \frac{ g_w^2 g_s^2 N_F D6}{6} G^a_{ij} G^a_{ij} A_0^2 \, .
\end{eqnarray}
There are also pure glue terms which are identical to the pure gauge
terms listed in the text, but with the substitution $N_d \rightarrow
2 N_f$ where $N_f = 6$ is the number of quark flavors. The 2 is because
both right and left handed quarks couple to the gluons, wereas the
weak coupling is chiral.
|
\section{Introduction}
The Minimal Supersymmetric Standard Model\cite{MSSM} (MSSM) is one of the
leading candidate theories for physics beyond the Standard Model (SM). The MSSM
is a globally supersymmetric version of the SM, where supersymmetry
breaking is implemented by the explicit introduction of soft-supersymmetry
breaking terms. The MSSM is minimal in the sense that the fewest number of
additional new particles and interactions are incorporated which are
consistent with phenomenology.
In particular, possible baryon ($B$) and lepton ($L$)
number violating interactions are excluded from the superpotential
(the presence of {\it both} $B$ and $L$ violating interactions can lead to
catastrophic proton decay rates).
As a result, there exists a
conserved $R$-parity, where the multiplicative quantum number $R=+1$ for
ordinary particles, and $R=-1$ for superpartners.
A consequence of $R$-parity
conservation is that the lightest supersymmetric particle (LSP) is
absolutely stable.
Theoretical prejudice coupled with experimental constraints
strongly favor a color and charge neutral LSP\cite{WOLF}.
In addition, in the MSSM, the
LSP is strongly favored to be the massive, weakly interacting
lightest neutralino $\widetilde Z_1$\cite{BDT,JELLIS}.
$\widetilde Z_1$'s, if they exist, would have been abundantly produced
in the early universe; if so, then relic neutralinos could well
make up the bulk of the dark matter in the universe today\cite{KT,JKG}.
In this paper, we restrict ourselves to the framework of the low energy
effective Lagrangian which is expected to result from, for instance,
supergravity grand unified models\cite{ARN}.
In these models, it is assumed that supersymmetry is broken in a hidden sector
of the theory. Supersymmetry breaking is then communicated to the
observable sector via gravitational interactions, leading to a common
mass $m_0$ for all scalar particles, a common mass $m_{1/2}$ for all
gauginos, a common trilinear coupling $A_0$, and a bilinear coupling $B_0$.
These soft supersymmetry breaking parameters are induced at energy scales
at or beyond the unification scales, but with (theoretically motivated)
values typically in the range $100-1000$ GeV.
The resulting theory, the MSSM with universal soft breaking terms,
is then regarded as an effective
theory with Lagrangian parameters renormalized at
an ultra-high scale $M_X \sim M_{GUT}-M_{Planck}$, and valid only
below this scale.
The corresponding weak scale sparticle couplings and masses can
then be calculated by evolving 26 parameters via
renormalization group equations\cite{RGE}
from the unification scale to the weak scale. An elegant
by-product\cite{RAD} of this mechanism
is that one of the Higgs boson mass squared terms is driven negative,
resulting in a breakdown of electroweak symmetry.
The radiative electroweak symmetry breaking constraint allows one to
essentially eliminate $B$ in favor of $\tan\beta$ (the ratio of Higgs field
vev's), and to calculate the magnitude of the superpotential Higgs mixing
term $\mu$ in terms of $M_Z$ (where we actually minimize the full one-loop
effective potential).
The model is then specified by only four
SUSY parameters (in addition to SM masses and couplings).
A hybrid set consisting of the common GUT scale scalar mass $m_0$,
common gaugino mass $m_{1/2}$, common SUSY-breaking trilinear
coupling $A_0$, along with the weak scale value of $\tan\beta$
proves to be a convenient choice. In addition, the sign of $\mu$ must
be stipulated.
These parameters fix the weak scale masses and couplings of all the
sparticles\cite{SPECTRA}.
The matter density of the universe $\rho$ is usually
parametrized\cite{KT,JKG} in terms of $\Omega=\rho /\rho_c$,
where $\rho_c ={3H_0^2/8\pi G_N} \simeq 1.88\times 10^{-29}h^2$
g/cm$^3$, and $h$, the Hubble scaling constant, is related to the Hubble
constant $H_0$ by $H_0=100h$ km/sec/Mpc. Here $h$ parametrizes our ignorance
of the true value of $H_0$, so that $0.5\alt h\alt 0.8$. Measurements of
galactic
rotation curves suggest $\Omega\sim 0.03 - 0.1$, compared to a luminous matter
density of $\Omega_{lum.}\alt 0.01$. Galactic clustering and galactic flows
suggest even larger values of $\Omega\sim 0.2-1$. Finally, the
theoretically attractive inflationary cosmological models require a flat
universe with $\Omega=1$. Meanwhile, estimates of the baryonic contribution
to the matter density of the universe from Big-Bang nucleosynthesis
suggest that $\Omega_{baryonic}\sim 0.01-0.1$. These analyses and estimates
suggest that the bulk of matter in the universe is (non-baryonic) dark matter.
Finally, analyses of structure formation in the universe in light of the COBE
measurements of anisotropies in the cosmic microwave background radiation
suggest that dark matter may be made of $\sim 60\%$ cold-dark matter
(weakly interacting massive particles or WIMPS, such as the lightest neutralino
$\widetilde Z_1$), $\sim 30\%$ hot dark matter (such as relic neutrinos), and
$\sim 10\%$ baryonic matter. This is the so-called
``mixed dark matter '' scenario.
The central idea\cite{KT} behind relic density calculations is that in the very
early
universe, neutralinos were being created and annihilated, but that they
were in a state of thermal equilibrium with the cosmic soup. As the universe
expanded and cooled, temperatures dropped low enough that neutralinos could
no longer be produced ($T\alt m_{\widetilde Z_1}$),
although they could still annihilate with one
another, at a rate governed by the total neutralino pair annihilation
cross section, and the neutralino number density.
Ultimately, as the universe expanded further, the expansion rate
outstripped the annihilation rate, thus freezing out the remaining neutralino
population of the universe, and locking in a neutralino relic density.
Our goal in this paper is to carry out estimates of the neutralino relic
density expected from the minimal SUGRA model. One solid constraint on
supersymmetric models with relic dark matter particles comes from the age
of the universe, which ought to be greater than $10$ ($15$) Gyrs; this
implies $\Omega h^2< 1$ ($0.25$). Thus, models with too large a relic
density would yield too young of a universe, in violation at least
with the age of the oldest stars in globular clusters. Furthermore,
models with $\Omega h^2< 0.025$ would not even be able to account for the dark
matter needed to explain galactic rotation: such models would be considered
cosmologically uninteresting. Models with intermediate values of
$0.025\alt \Omega h^2\alt 1$ are considered cosmologically interesting,
as they might explain galactic rotation and clustering, or might even
make up the matter density needed for inflationary cosmology, given a cold-dark
matter (CDM: $\Omega h^2\sim 0.25-0.64$) or mixed hot/cold dark matter
scenario (MDM: $\Omega h^2\sim .15-.4$).
Following the procedures outlined by Lee and Weinberg\cite{LW}, many groups
have calculated the relic neutralino
abundance\cite{HAIM,ELLIS,GRIEST,OLIVE,DREES,LOPEZ,ROSZ,AN}.
Early works involved calculating the most important
neutralino annihilation channels, usually assuming the LSP was a
photino. Later studies included various improvements, including
more annihilation channels, more general neutralino mixings, and more
realistic supersymmetric particle spectra.
A common thread amongst many papers was the calculation of
the Boltzmann-averaged
quantity $\sigma\times v$ using a power series expansion in
velocity. Such an approach was shown to be inaccurate when relativistic
effects were important, when annihilation proceeded through $s$-channel
poles, when threshold effects were important, or when
co-annihilation processes occured\cite{GS,GG}.
Many recent calculations have included some or all of these effects.
We have several goals in mind for the present paper.
\begin{itemize}
\item We wish to present reliable calculations for the neutralino relic
density in supersymmetric models. To this end we evaluate {\it all} $2\rightarrow 2$
neutralino annihilation diagram amplitudes numerically as complex numbers,
without approximation.
We perform Boltzmann averaging using the Gondolo-Gelmini formalism\cite{GG}.
This takes into account relativistic thermal averaging, while our numerical
helicity amplitude technique avoids the usual uncertainties inherent in
the velocity expansion, so that Breit-Wigner poles and threshold effects
are fully accounted for. Co-annihilation can occur when the two lightest
superpartners are very close in mass-- this situation rarely occurs within the
SUGRA framework adopted in this paper, and hence we ignore it.
\item We present results in the well-motivated SUGRA framework, which
includes gauge coupling unification, Higgs mass radiative
corrections\cite{HIGGS},
and radiative electroweak symmetry breaking using the one-loop effective
potential\cite{GRZ}.
\item Our results for the relic density calculation can be directly
related to recent calculations for various supersymmetry signals
expected at the LEP2\cite{LOPLEP,LEP2}, Tevatron\cite{BKT,LOPTEV,BCKT,MRENNA}
and LHC colliders\cite{BCPT,CHEN}.
In particular,
relic density calculations have a preference for light sleptons.
Such light sleptons may well be observable at LEP2 or LHC colliders, and
yield enhanced rates for $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$ states at the Tevatron
collider\cite{BT}.
\end{itemize}
To accomodate these goals, we present in Sec. II various details of our
relic density calculation, including those peculiar to the present approach.
In Sec. III, we present numerical results for our relic density calculations
in the $m_0\ vs.\ m_{1/2}$ plane of the minimal SUGRA model. In Sec. IV,
these results are explicitly compared to expectations for minimal
SUGRA at various
collider experiments, as worked out in a series of previous papers.
Finally, in Sec. V we present an overview and some conclusions.
\section{Calculational details}
We begin our determination of the neutralino relic density from minimal
supergravity by selecting a point in the SUGRA parameter space
\begin{eqnarray}
m_0,\ m_{1/2},\ \tan\beta ,\ A_0\ {\rm and}\ sign(\mu ),
\end{eqnarray}
where in addition we take the top quark mass $m_t=170$ GeV. The 26
renormalization group equations are iteratively run between the weak scale
and the GUT scale, which is defined as the point where the $U(1)$, $SU(2)$
and $SU(3)$ gauge couplings unify, and is typically $M_X\sim 2\times 10^{16}$
GeV. We use 2-loop RGE equations for gauge and Yukawa couplings
(with SUSY particle threshold effects), but only
1-loop equations for the running of the various soft breaking terms. The
1-loop effective Higgs potential is minimized to enforce radiative electroweak
symmetry breaking. Our procedure has been described in more detail in
Ref. \cite{BCMPT}, and has been incorporated into the event generator
ISAJET\cite{ISAJET}. At this point, a correlated sparticle mass spectrum
and couplings emerge from our input point in SUGRA parameter space.
The next step in our computation, after obtaining the superparticle spectrum,
is to evaluate the neutralino relic density by solving the Boltzmann
equation as formulated for a Friedmann-Robertson-Walker
cosmology\cite{KT}. Central to the evaluation of the relic density is the
computation of the fully relativistic,
thermally averaged neutralino annihilation cross section
times velocity, defined as
\begin{eqnarray}
<\sigma v_{Mol}>(T)={
{\int \sigma v_{Mol}e^{-E_1/T} e^{-E_2/T} d^3p_1 d^3p_2}\over
{\int e^{-E_1/T} e^{-E_2/T} d^3p_1 d^3p_2}
},
\end{eqnarray}
where $p_1$ ($E_1$) and $p_2$ ($E_2$) are the momentum and energy of the two
colliding particles in the cosmic, co-moving frame of reference, and $T$
is the temperature.
The above expression has been reduced to a one-dimensional integral by
Gondolo and Gelmini\cite{GG}, which yields
\begin{eqnarray}
<\sigma v_{Mol}>(x)={1\over{4xK_2^2({1\over x})}}
\int_2^{\infty} da \sigma (a) a^2(a^2-4) K_1({a\over x}),
\end{eqnarray}
where $x={T\over m_{\widetilde Z_1}}$, $a={\sqrt{s}\over m_{\widetilde Z_1}}$, $\sqrt{s}$ is
the subprocess energy, and $K_i$ are modified Bessel functions of order $i$.
We evaluate the neutralino annihilation cross section for
$\widetilde Z_1\widetilde Z_1\rightarrow f_1 f_2$ as
\begin{eqnarray}
d\sigma (a)={1\over {32\pi s}}
{{\lambda^{1\over 2}(s,m_{f_1}^2,m_{f_2}^2)}\over
{\lambda^{1\over 2}(s,m_{\widetilde Z_1}^2,m_{\widetilde Z_1}^2)}}
{\overline\Sigma} |{\cal M}|^2 d\cos\theta ,
\end{eqnarray}
where ${\overline\Sigma}|{\cal M}|^2$ is the spin summed and averaged
squared matrix element. Our calculation of the relic density is distinct in
that we evaluate ${\cal M}$ for {\it all} Feynman diagrams listed in Table 1
as complex numbers, using the HELAS\cite{HELAS} helicity amplitude
subroutine package. Thus, our approach avoids the usual uncertainties
associated with the expansion of cross section in terms of a power series in
velocity. The integration over $\cos\theta$ is performed numerically using
Gaussian quadratures.
To evaluate the neutralino relic density, the freeze out temperature $x_F$
is needed. The standard procedure here to iteratively solve the
freeze out relation
\begin{eqnarray}
x_F^{-1}=\log \Big[ {m_{\widetilde Z_1}\over {2\pi^3}} {\sqrt{45\over {2g_* G_N}}}
<\sigma v_{Mol}>_{x_F} x_F^{1\over 2}\Big],
\end{eqnarray}
by starting with a trial value $x_F={1\over 20}$. In the above, $g_*$ is the
effective number of degrees of freedom at $T=T_F$ ($\sqrt{g_*}\simeq 9$),
and $G_N$ is Newton's constant.
Finally, the relic density can be calculated from
\begin{eqnarray}
\Omega h^2= {\rho (T_0)\over {8.0992\times 10^{-47}\ {\rm GeV}^4}},
\end{eqnarray}
where
\begin{eqnarray}
\rho (T_0)\simeq 1.66\times{1\over M_{Pl}}
({{T_{m_{\widetilde Z_1}}}\over{T_\gamma}})^3
T_\gamma^3 {\sqrt{g_*}} {1\over {\int_0^{x_F}<\sigma v_{Mol}> dx}}.
\end{eqnarray}
To evaluate the integral in the above expression, we expand the modified
Bessel functions in Eq. 2.3 as power series in $x$, and then integrate over
$x$.
The result is
\begin{eqnarray}
\int_0^{x_F}<\sigma v_{Mol}> dx ={1\over {8\pi}} \int_2^\infty da \sigma (a)
a^{3\over 2} (a^2-4) F(a),
\end{eqnarray}
where
\begin{eqnarray*}
F(a)&=&\sqrt{{\pi\over {a-2}}} \left\{ 1-Erf(\sqrt{{{a-2}\over
x_F}})\right\}+\\
& &2({3\over 8a}-{15\over 4})\left\{ \sqrt{x_F} e^{-{{a-2}\over x_F}}-
\sqrt{\pi (a-2)} (1-Erf(\sqrt{{{a-2}\over x_F}}))\right\}+ \\
& & {2\over 3}({285\over 32}-{45\over 32a}-{15\over 18a^2})\times\\
& &\left\{ e^{-{{a-2}\over x_F}}\left[ x_F^{3\over 2}-2(a-2)\sqrt{x_F}\right]
+2\sqrt{\pi }(a-2)^{3\over 2}(1-Erf(\sqrt{{a-2}\over x_F} ))\right\} .
\end{eqnarray*}
In the above, virtually all the contribution to the integral comes from
$x<2.5$. We integrate the above expression numerically with Gaussian
quadratures, taking care to scan finely the regions with a Breit-Wigner pole.
In the region of a pole, the domain of integration must be broken into very
tiny
intervals, and obtaining convergence for a single point in parameter space
can take up to several hours of CPU time on a DEC ALPHA.
\section{Results from relic density calculation}
Our first numerical results for the relic density from minimal SUGRA models
are given in Fig. 1, where we plot contours of the neutralino relic
density $\Omega h^2$ in the $m_0\ vs.\ m_{1/2}$ parameter plane, where
we take $A_0=0$, $\tan\beta =2$, $\mu <0$ and $m_t=170$ GeV.
Changes in the $A_0$ parameter mainly affect 3rd generation sparticle
masses, and consequently result in only small changes in the relic density.
The regions labelled TH are excluded by theoretical considerations:
either there is a charged or colored LSP (or the $\tilde\nu$ is LSP), or
the radiative electroweak symmetry breaking constraint breaks down.
The region labelled by EX corresponds to parameter space already excluded
by SUSY searches at LEP or Fermilab Tevatron experiments\cite{BCMPT}.
In almost all of
the plane, we find $\Omega h^2 >0.025$, {\it i.e.} large enough to explain the
galactic rotation curves. However, the region to the right of the
$\Omega h^2 =1$ contour is certainly excluded in that the age of the
universe would be younger than $10$ Gyrs. Meanwhile, a dominantly CDM
inflationary universe would lie in between the $\Omega h^2 =0.25-0.75$
contours. The COBE favored MDM inflationary universe would lie
between the $\Omega h^2 =0.15-0.4$ contours. For this latter favored region,
$m_{1/2}$ is bounded by $m_{1/2}\alt 400$ GeV (corresponding to
$m_{\tilde g}\alt 1000$ GeV), and $m_0<150$ GeV, unless the gluino is very
light ($m_{\tilde g}\simeq 300$ GeV).
(For comparison, various SUSY particle mass contours
for the same parameter choices are listed in Refs. \cite{BCMPT,BCKT,BCPT}.)
We find in general that large values of $m_0\agt 350$ GeV
(corresponding to $m_{\tilde\ell}\agt 250$ GeV) yield too young
a universe (due to suppression of $t$-channel slepton exchange diagrams),
except for the two narrow corridors in the lower right region
of the figure. The upper of the two corridors corresponds to neutralino
annihilation through the $Z$ pole, so the relic density is largely reduced
by $Z$ mediated $s$-channel annihilation diagrams. The lower of the two
corridors corresponds to annihilation through an $s$-channel light Higgs
$h$ pole-- in this case, the relic density falls rapidly to values even
below $\Omega h^2\sim 0.025$.
A qualitative feel for the relative importance of different annihilation
channels can be gleaned from Fig. 2. Here we plot for $m_0$ fixed at
200 GeV, as a function of $m_{1/2}$, the thermally averaged
annihilation cross section times velocity, integrated over temperature,
which enters into the relic density calculation (Eq. 2.8).
Larger cross sections correspond to smaller relic densities. As $m_{1/2}$
increases, the first pole we come to is annihilation via $s$-channel $h$,
where $\widetilde Z_1\widetilde Z_1\rightarrow b\bar b$ dominates. In these plots, $m_{\widetilde Z_1}$ scales
with $m_{1/2}$, and at the Higgs pole in this plot (on the edge of exclusion
by LEP Higgs search experiments),
$m_{\widetilde Z_1}\simeq 30$ GeV and $m_{\widetilde W_1}\simeq 70$ GeV.
As one moves to higher $m_{1/2}$,
annihilation through the $Z$ pole is reached, which is dominated by
$\widetilde Z_1\widetilde Z_1\rightarrow d\bar d,\ s\bar s$, and $b\bar b$.
For values of $m_{1/2}$ away from poles,
annihilation via $t$-channel slepton and sneutrino exchange dominates.
For even higher
$m_{1/2}$ values, annihilation into channels such as $hh,\ Zh,\ WW$ and $ZZ$
open up, but never dominate for the parameter choices in this plot.
Annihilation into other channels such as $HA$, $AA$, $HH$ amd $H^+H^-$
are included in our calculation, but unimportant given our SUGRA sparticle
mass spectrum, which yields very large masses for Higgs bosons other than $h$.
The onset of the $\widetilde Z_1\widetilde Z_1\rightarrow t\bar t$ can be detected in the
$\widetilde Z_1\widetilde Z_1\rightarrow\Sigma u_i\bar{u_i}$ curve around $m_{1/2}\sim 400$ GeV.
If we plot the relic density for the same parameter choices, but flip the
sign of $\mu$, so that $\mu >0$, then we obtain the results of Fig. 3. The
relic density contours in this case are similar to those of Fig. 1 for
large values of $m_{1/2}$, where annihilation dominantly occurs via
slepton exchange. The kink in the contours is due to the onset of the
$\widetilde Z_1\widetilde Z_1\rightarrow t\bar t$ channel.
In this case, annihilation through $t$-channel $\tilde t_1$ exchange makes
a large contribution to the total annihilation cross section.
For smaller values of $m_{1/2}$,
in contrast to Fig. 1, we find only one
corridor extending to large $m_0$ where the relic density drops to
cosmologically un-interesting values. In this case, the $Z$ and $h$ poles
very nearly overlap for $m_{\widetilde Z_1}\sim 46$ GeV. This can be seen in more detail
in Fig. 4, where again we show the thermally averaged cross section versus
$m_{1/2}$, for $m_0 =200$ GeV.
Finally, we show again the neutralino relic density $\Omega h^2$ in the
$m_0\ vs.\ m_{1/2}$ plane for the same parameter choices as Fig. 1, except
now we take a large value of $\tan\beta =10$. For this case, we note the
rather broad band at $m_{1/2}\sim 100-140$ GeV, where $\Omega h^2 <0.025$-
too low to explain even the galactic rotation curves, and due again to
annihilation through the $s$-channel graphs. In fact, inflationary models,
which require $\Omega h^2 \agt 0.15$, are only allowed if $m_{1/2}>150$ GeV,
corresponding to $m_{\tilde g}>400$ GeV. In this plot, there is a significant
region extending to large values of $m_0$, corresponding to large $m_{\tilde q}$
and large $m_{\tilde\ell}$, for $m_{1/2}\sim 150-190$ GeV. The contributing
thermally averaged subprocess cross sections are again shown in Fig.~6 for
$m_0=200$ GeV. In this plot, the $Z$ pole annihilation channel occurs at
$m_{1/2}\simeq 110$ GeV, followed by the Higgs pole at $m_{1/2}\simeq 130$ GeV.
The rough overlap of these two pole contributions leads to the single broad
corridor of low $\Omega h^2$ shown in Fig. 5.
\section{Implications for SUSY searches at colliders}
Recently, various papers have been written on the prospects for
supersymmetry at the LEP2 $e^+e^-$ collider\cite{LOPLEP,LEP2}, the Fermilab
Tevatron $p\bar p$ collider\cite{BT,BKT,LOPTEV,BCKT,MRENNA} and
the CERN LHC $pp$ collider\cite{BCPT,CHEN}.
Our objective in this section is to assess the prospects
for discovery of SUGRA at hadron and $e^+e^-$ colliders, given the additional
constraints from requiring a reasonable value for the neutralino relic
density. We mainly focus on
the collider results of Refs. \cite{LEP2,BKT,BCKT,BCPT,CHEN,DPF},
since they were performed
in a consistent framework, in the same $m_0\ vs.\ m_{1/2}$ plane.
In Fig. 7, we again show the neutralino relic density contours in the
$m_0\ vs.\ m_{1/2}$ parameter plane, for the same parameter choices as in
Fig. 1. In addition, we have added on contours for SUSY discovery at various
colliders. Supersymmetric particles ought to be discoverable at LEP2
operating at $\sqrt{s}=190$ GeV, with integrated luminosity
$\int {\cal L}dt=300$ fb$^{-1}$ below the contour labelled LEP2\cite{LEP2}.
The
lower-left bulge in the LEP2 contour is where sleptons ought to be detectable,
while beneath the contour running along $m_{1/2}\simeq 100$ GeV (which
runs through the neutralino $Z$-pole annihilation region), charginos
ought to be detectable. By comparing, we see that the region accessible by LEP2
generally has $\Omega h^2<0.15$ {\it i.e.} not the most cosmologically
favored region, but with enough dark matter to explain galactic rotation.
However, the contour labelled with LEP2-Higgs shows the reach of LEP2 for
the light SUSY Higgs boson $h$, which is just below $m_{1/2}\sim 400$ GeV.
This region completely encloses the favored MDM region. The implication is
that if MDM explains dark matter in the universe, and if $\tan\beta$ is small
and $\mu <0$, then LEP2 ought to discover at least the light SUSY Higgs boson.
The dashed contour labelled Tevatron is a composite of the reach of
Tevatron Main Injector era ($\sqrt{s}=2$ TeV; $\int {\cal L}dt=1$ fb$^{-1}$
integrated luminosity) experiments for multi-jet$+E\llap/_T$ events\cite{BKT},
and mainly, for $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$ events\cite{BCKT}. We see that the
largest reach by Tevatron experiments occurs exactly in the
cosmologically favored MDM region, and can reach to $m_{1/2}\sim 160$ GeV,
corresponding to $m_{\tilde g}\sim 440$ GeV. This is no accident: a reasonable
neutralino annihilation cross section generally requires $m_{\tilde\ell}\alt 200$
GeV; these lighter sleptons give rise to enhanced leptonic decay of
neutralinos,
leading to large rates for $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$ events.
Since lower values of $m_{1/2}$
are preferred by fine-tuning arguments\cite{FT}, there is a good chance
Tevatron experiments could discover SUSY via $3\ell$ events if nature chose
this parameter set.
We also compare the results of Fig. 7 with expectations for supersymmetry
at the CERN LHC collider. Of course, LHC experiments can cover the whole
parameter plane up to $m_{1/2}\sim 600-800$ GeV with only $\int {\cal L}dt=10$
fb$^{-1}$ of integrated luminosity, at $\sqrt{s}=14$ TeV, via searches for
multi-jet$+E\llap/_T$ events from gluino and squark cascade decays\cite{BCPT},
so discovery of SUSY would be no problem.
We also plot in Fig. 7 the contour beneath which
sleptons ought to be visible at LHC\cite{BCPT,CHEN}. We see that the
cosmologically favored MDM region falls almost entirely within the slepton
discovery region, so that if the MDM scenario is correct, then LHC has a very
high
probability to discover a slepton.
Since sleptons are relatively light, LHC experiments ought as well to be
sensitive to $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$ events over much, but not all, of the
favored MDM region\cite{BCPT,CHEN}.
(In some of the favored region, $\widetilde Z_2\rightarrow \nu\tilde\nu$ or
$\widetilde Z_1 h$, thus spoiling the signal.)
Finally, since $m_{\tilde\ell}\alt 250$ GeV in the MDM scenario, sleptons would
then likely be visible at a linear $e^+e^-$ collider operating
at $\sqrt{s}=500$ GeV.
In Fig. 8, we show the same relic density contours as in Fig. 3 ($\tan\beta =2$
, $\mu >0$), and compare again with expectations for colliders. In this case,
we see the LEP2 contour again lies in a region of $\Omega h^2 <0.15$,
although it does encompass the cosmologically interesting region around
$(m_0,m_{1/2})\sim (100,110)$. The LEP2 Higgs contour in this case lies at
$m_{1/2}\sim 170$ GeV, and thus covers only a portion of the MDM favored
region.
Thus, if the MDM scenario is correct, and $\tan\beta$ is small, minimal SUGRA
sparticles or light Higgs boson might still not be accessible at LEP2.
We also plot the contour due to the combined Tevatron MI reach. In this case,
there is a large Tevatron reach due to $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$
extending to $m_{1/2}\sim 230$ GeV, overlapping considerably with
the MDM region. Finally, we note once again that LHC can cover the whole plane
via multi-jet$+E\llap/_T$ searches. In addition, the MDM region lies again almost
entirely within the LHC slepton search region, and overlaps substantially
with the LHC $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$ clean trilepton region\cite{CHEN}.
Last of all, we turn to Fig. 9, which compares the neutralino relic
density with collider search regions for large $\tan\beta =10$, with $\mu <0$.
In this case, we note that the MDM favored region lies entirely above
the region that is searchable at LEP2. In addition, for this case, the lightest
Higgs boson has mass $m_h \agt 90$ GeV throughout the plane, beyond the
reach of LEP2 at $\sqrt{s}=190$ GeV. Hence, if $\tan\beta$ is large, and the
MDM scenario is correct, then there would be little hope of seeing SUSY at
LEP2.
In this case, the prospect for minimal SUGRA at Tevatron MI is even worse,
except for the narrow region extending along $m_0\sim 100$ GeV, which enters
into the cosmologically favored MDM region. Finally, we note that once again
the LHC slepton reach contour excloses most of the MDM region, with the main
exception being the band of allowed MDM region extending to large $m_0$
along $m_{1/2}\sim 160-170$ GeV. The LHC $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$ region
encloses pieces of the MDM region, but leaves significant
areas uncovered\cite{CHEN}.
\section{Conclusion}
In this paper, working within the minimal supergravity model with
radiative electroweak symmetry breaking and universal GUT scale
soft supersymmetry breaking terms, we have evaluated the cosmological
relic density from neutralinos produced in the early universe. Our technique
was to evaluate {\it all} lowest order neutralino annihilation Feynman
diagrams as complex helicity amplitudes. We then performed the necessary
integrations numerically, preserving relativistic covariance, and avoiding the
usual expansion as a power series in velocity. While this approach might be
regarded as a brute force numerical calculation, it does include
relativistic thermal averaging, annihilation threshold effects, and careful
integration over Breit-Wigner poles. We do not include co-annihilation
processes in our calculations, which are however unimportant within the SUGRA
framework, in which we work.
Our numerical results for the neutralino relic density were presented in
Figures 1-6. For the favored mixed dark matter scenario, for which
$0.15<\Omega h^2 <0.4$, we find that, unless annihilation occurs via
$s$-channel
$Z$ or $h$ exchange (in which case $m_{\tilde g}<300-400$ GeV\cite{BDKNT}),
$m_{\tilde g}\alt 1000$ GeV, and $m_{\tilde\ell}\alt 250$ GeV. The less conservative
constraint from the age of the universe ($\Omega h^2 <1$) yields larger bounds
on sparticle masses.
We also examined the implications of our relic density calculations for
collider searches for the sparticles of minimal SUGRA. These results have
been summarized in Figs. 7-9. Within the MDM range of $\Omega h^2$, we find
that
LEP2 has a high probability to detect a light Higgs boson if
$\tan\beta$ is small and $\mu <0$. For the opposite sign of $\mu$, $m_h$ can be
larger, and detection at LEP2 is less certain. Prospects for detection
of sleptons or charginos at LEP2 are less bright: generally, if $m_{\tilde\ell}<90$
GeV, $t$-channel neutralino annihilation is too large, leading to rather
low values of neutralino relic density. Likewise, if $m_{\widetilde W_1}<90$ GeV, then
$m_{\widetilde Z_1}\alt 45$ GeV, and neutralinos can annihilate via $s$-channel $Z$
or $h$ exchange, again leading to only a small relic abundance.
Prospects for discovering SUGRA at Tevatron MI experiments are
somewhat brighter, since a reasonable relic density requires roughly
$100< m_{\tilde\ell}<250$ GeV. Such a slepton mass range generally leads to enhanced
leptonic decays of neutralinos, giving Tevatron experiments a good chance to
find SUGRA via $\widetilde W_1\widetilde Z_2\rightarrow 3\ell$ searches.
The CERN LHC $pp$ collider can make a thorough search for supersymmetry
over all the allowed parameter space in the multi-jet $+E\llap/_T$ channel.
However, the rather light slepton masses required for reasonable neutralino
relic densities falls within the range of LHC experimental sensitivity, so
there is a good chance to find sleptons at LHC if, for instance, the MDM
scenario turns out to be correct. Likewise, experiments at
an $e^+e^-$ linear collider
operating at $\sqrt{s}\sim 500$ GeV would stand a good chance of discovering
sleptons, since they would be sensitive to slepton pair production for
$m_{\tilde\ell}\alt 230$ GeV\cite{HIT,DPF}.
\smallskip
\noindent{\it Note added: Upon completion of this work, a preprint
appeared which addressed the neutralino dark matter relic density in
SUGRA models with non-universal soft-breaking terms\cite{NEWELLIS}.}
\acknowledgments
We thank X. Tata and C. H. Chen for discussions, and X. Tata
for comments on the manuscript.
This research was supported in part by the U.~S. Department of Energy
under grant number DE-FG-05-87ER40319.
|
\section{Introduction}
\label{Introduction}
A two--channel Kondo impurity in a metal is, perhaps, the most promising
impurity system to exhibit low temperature non--Fermi--liquid behavior
experimentally. This hope rests on the relatively few degrees of
freedom involved (a local spin doublet coupled to two degenerate
``flavors'' of spin--$\frac{1}{2}$ fermions) with the symmetry group
under which they transform being not very complicated
[$SU(2)_{flavor}\times SU(2)_{spin}\times U(1)_{charge}$]. As a result
one might hope that it would not be difficult to find a system
whose low temperature behavior would be described by the two--channel
Kondo system.
Yet, for more than a decade since it was first
introduced by Nozi\'{e}res and Blandin\cite{Nozieres1}, no
experimental realization of this model has been conclusively
demonstrated. The difficulty lies in the fact that the
non--Fermi--liquid fixed point is unstable to various
symmetry--breaking processes which turn out to be present in real
experimental situations. For example, Nozi\'{e}res
and Blandin\cite{Nozieres1} pointed out that anisotropy between
the two flavor channels, caused by lattice effects, would destroy the
non--Fermi--liquid ground state. Since this anisotropy could not be
made to vanish for the known cases, the search for
non--Fermi--liquid behavior with conventional spin--Kondo systems was
suspended. Cox\cite{Cox} pointed out, however, that under certain
symmetry conditions, local quadrupolar degrees of freedom could
result in two channel Kondo-like coupling; unfortunately, dilute
impurity systems of this type have proven hard to make.
Later, Vlad\'{a}r and Zawadowski\cite{Zawadowski1} suggested that a
non--magnetic impurity tunnelling between two sites in a metal could
be modeled as a two--channel Kondo system in which the roles of the
channels and the spins in the original formulation are
interchanged. In this system the spin of the electron plays the role
of the ``flavor channels'', so that the anisotropy between
``channels'' is no longer an issue
since in zero external magnetic field the spin--up and the
spin--down electrons are degenerate. This led Vlad\'{a}r and
Zawadowski\cite{Zawadowski1} to predict
non--Fermi--liquid behavior in such a system. Recently Ralph
{\em et al}\cite{Ralph1,Ralph2} have interpreted low
temperature tunnelling data in very small metallic contacts in terms
of two--channel Kondo--like physics. Their measurements are
claimed to be consistent with certain exact results obtained by Affleck and
Ludwig\cite{Affleck1}, but at this point, the interpretation is
still controversial\cite{Wingreen}.
Unfortunately, the mapping of the two--site impurity to the
two--channel Kondo (2CK) system is far from exact, and, even with
no anisotropy between ``channels'' (the spin--up
and spin--down electrons) there are other processes present in the tunnelling
impurity system which are relevant and hence generically destroy the
non--Fermi--liquid ground state. These processes, which cannot be
neglected have not been treated adequately in the
literature\cite{Zawadowski1,Zimanyi1,Murumatsu}.
In this paper we carefully consider the mapping between the tunnelling
impurity system and the two--channel Kondo (2CK) problem. We first
analyze the behavior when the impurity tunnelling starts to become
important as the temperature is lowered; this is analogous to the
weak coupling regime of the Kondo problem and is needed in order to
understand which tunnelling processes will dominate at low
temperatures. We then analyze the intermediate coupling behavior to
investigate whether there are parameter regimes in which the system
will be governed by the 2CK fixed point over a reasonable range of
temperatures -- as would be needed to observe the non--Fermi liquid
behavior experimentally.
Unfortunately, even in the optimal case of an impurity tunnelling
between two identical sites so that the system has an extra $Z_2$
symmetry, we find that the 2CK fixed point is only accessible if two
tunnelling processes, which are very different physically, nearly
exactly cancel. Generically, the system will exhibit Fermi liquid
behavior at low temperatures.
We analyze the behavior near the 2CK point in detail,
focusing on the connection between the physical operators and those
that appear as ``natural'' operators in the 2CK language. It is found
that, on the critical manifold which can be obtained in the symmetric
impurity problem by adjusting one parameter, there are {\em four}
leading irrelevant directions in contrast to the behavior for the pure
2CK problem analyzed by Sengupta and Georges\cite{Georges}.
Various symmetry breaking terms are also studied and our results
recover those derived by Affleck {\em et al}\cite{Affleck2,Affleck3} with
conformal field theory techniques.
A somewhat surprising feature emerges in the fuller phase diagram of
the symmetric impurity model: a {\em second} fixed point which
exhibits non-Fermi liquid behavior, albeit one with two
relevant directions in the $Z_2$ symmetric case.
\subsection{Outline}
In the remainder of the Introduction we motivate the form of the
Hamiltonian in which we focus and interpret the various terms that
should appear. In Section II we derive an effective Hamiltonian that
we will study and analyze its symmetries, while in Section III the
weak coupling analysis is outlined. In Section IV the behavior near
the intermediate coupling 2CK fixed point and, for completeness, the
various symmetry breaking operators near the 2CK fixed point are
studied. In Section V we discuss the existence of an extra novel fixed
point at intermediate coupling. In Section VI, we discuss the
accessibility of the 2CK fixed point and draw our conclusions.
Finally, in the Appendices, the details of the weak--coupling
(Appendix A) and the intermediate--coupling analysis (Appendix B) are
presented; the comparison of our results to those obtained by
conformal field theory is made in Appendix C.
\subsection{Physical Picture}
The system we wish to describe is an impurity or heavy particle which
can hop back and forth between two sites coupled to a bath of
electrons. The two sites may or may not be equivalent but we will
primarily focus on the symmetric (equivalent) case. The asymmetric
case can readily be treated in a similar manner. The effects of the
interaction of this impurity with the electrons can be manifested in a
number of ways. First, the electrons will tend to screen the charge of
the impurity. Thus the impurity will hop between the two sites
carrying with it tightly bound electrons which can move fast enough to
adjust to the position of the impurity; it is convenient to consider
these to simply be part of the ``dressed'' impurity particle. However,
in addition, as the impurity moves it may also redistribute the low
energy electronic excitations near the Fermi surface. Since we are
interested in the low energy physics, these processes must be treated
directly. For simplicity we consider only $s$-wave scattering off the
impurity.
If an impurity of charge $Ze$ hops between two well separated sites,
then the Friedel sum rule relates the ($s$-wave) scattering phase
shift off a static impurity, $\delta$, to the electronic charge that
will be moved to screen the impurity as it moves adiabatically from
one site to the other, via $Z=2\delta/\pi$, with the factor of two due
to the two spin species. Conversely, if the two sites are close
together, one can still usefully speak of an effective charge $Q$ (per
spin) which plays an analogous role to $Q=\delta/\pi$ in the well
separated case, but is no longer simply related either to scattering
phase shifts or to the impurity charge. It will instead turn out to be
exactly the ``orthogonality catastrophe'' exponent that determines the
system size dependence of the overlap between the electronic ground
states with the impurity at the two sites.\cite{Yamada} For
simplicity, we will focus on $Q$ in the range $0\leq Q \leq 1$,
corresponding to repulsive interactions. (As shown in reference
\onlinecite{MF}, other ranges of the effective charge can be reduced
to this case via a set of more complicated combined impurity-electron
processes related to those we consider here.)
The important processes, in addition to hopping of the (dressed)
impurity by itself, will be those in which one or two low energy
electrons move in the {\em opposite} direction to that the impurity
hops. These processes can, for $Q>1/2$, reduce the effective charge
that must relax to the new impurity position, thereby decreasing the
orthogonality between the pre- and post-hop configuration; this
results in a larger amplitude for the combined impurity-electron
process at low temperatures, relative to the simple impurity
hop process. [Note that in reference \onlinecite{MF}, a
sign error in the definition of $Q$ resulted in the incorrect
interpretation being given to these processes; this error does not affect
the conclusions, just the interpretation of the combined hopping
processes].
In order to proceed with the analysis of the low temperature behavior
of interest it would appear to be important to assess the relative
magnitudes of the amplitudes, at some
high energy scale, of the processes discussed above. However, it has
been claimed\cite{Kagan} that this is
essentially impossible for strong electron-impurity
interactions. Indeed Kagan and Prokof'ev\cite{Kagan} have claimed that
a sensible Hamiltonian cannot be written in terms of a simple two
level system since the high energy electronic degrees of freedom
cannot be properly taken into account. Although there are indeed real
difficulties here, it should nevertheless be possible to introduce
{\em effective} amplitudes at some intermediate energy scale and then
analyze the behavior of the system in the phase space of these
effective parameters. In general, unless there are specific reasons to
prevent it, one would expect that most combinations of parameters
could, in principle, be realized. We thus approach the problem via
this route and start with an effective Hamiltonian at an intermediate
energy scale, at, say, some fraction of the conduction bandwidth. As
we shall see, the extra hopping processes will in any case be
generated at low energies.
\section{Effective Hamiltonian}
\label{Effective Hamiltonian}
The important electronic degrees of freedom at low energies are those
that interact with the impurity in one of its two positions ``1'' and
``2''. These are just the $s$-wave conduction electrons around the
positions of the two sites. Thus at each energy, there will be two
important electronic degrees of freedom per spin. However,
unless the two sites are very far apart (in which case the impurity
tunnelling rates will be negligible and hence not of interest), the
two sets of $s$-wave electrons will {\em not} be orthogonal; this will
play an important role in what follows. If we label the two sets of
$s$-wave electrons by their energy, $\epsilon$, measured from the
Fermi surface, then for each $\epsilon$ there is an (essentially)
unique pair of linear combinations of the two $s$-wave states, that
are orthonormal and transform into each other under interchange of the
two sites. We label the annihilation operators of this orthonormal
pair $c_{1\epsilon}$ and
$c_{2\epsilon}$ with anticommutation relations
\begin{equation}
\label{anticommutationscie}
\left\{ c_{i\epsilon},c^\dagger_{j\epsilon'}\right\} = 2\pi
\delta\left(\epsilon-\epsilon'\right) \delta_{ij}
\end{equation}
with the ``1'' and ``2'' denoting the sites near which the
wavefunction is larger. The impurity interacts with, simply, the
operators
\begin{equation}
\label{defcilocal}
c_{1,2}= \int \frac{d\epsilon}{2\pi} c_{1,2\epsilon},
\end{equation}
although in each position the impurity will couple to {\em both}
$c_1$ and $c_2$ due to the non-orthogonality of the two electronic
$s$-wave states. With time reversal invariance, which we assume
henceforth, the most general interaction with the impurity becomes
(ignoring spin for now)
\begin{eqnarray}
\label{Uinit}
U= d^\dagger_1d_1 \left[V_1 \left(c^\dagger_1c_1
+c^\dagger_2c_2\right)\right. &+& V_2
\left(c^\dagger_1c_2+c^\dagger_2c_1\right) \\ \nonumber
&+&V_3\left.\left(c^\dagger_1c_1
-c^\dagger_2c_2\right)\right] \\ \nonumber
+ d^\dagger_2d_2 \left[V_1 \left(c^\dagger_2c_2
+c^\dagger_1c_1\right) \right.&+& V_2
\left(c^\dagger_2c_1+c^\dagger_1c_2\right) \\ \nonumber
&+&V_3\left.\left(c^\dagger_2c_2
-c^\dagger_1c_1\right)\right]
\end{eqnarray}
where $d_{1,2}$ are the annihilation operator of the impurity at the sites.
Using the obvious identity $d^\dagger_1d_1+d^\dagger_2d_2=1$, Eq(\ref{Uinit})
can be written as
\begin{eqnarray}
\label{Ufinal}
U&=& V_1 \left(c^\dagger_1c_1 + c^\dagger_2c_2\right)
+ V_2 \left(c^\dagger_1c_2 + c^\dagger_2c_1\right) \\ \nonumber
&+& V_3 \left(d^\dagger_1d_1 -d^\dagger_2d_2\right) \left(c^\dagger_1c_1
-c^\dagger_2c_2\right).
\end{eqnarray}
Note the appearance of an effective electronic hopping term $V_2$,
caused by the scattering by the impurity; this will vanish if the
sites are far apart, but in general will be comparable to the other
terms.
The first term in Eq(\ref{Ufinal}) which is the average of the
interaction over the two impurity positions, merely produces a constant
phase shift for both ``1'' and``2'' electrons at the Fermi level and,
combined with other operators, gives rise only to irrelevant
terms. Therefore we will ignore it at this point
although in Section IV an irrelevant operator it
gives rise to will play a role in our discussion of the intermediate
coupling behavior.
With the effective hopping charge $Q$ in the range
$\left[0,1\right]$, there are three
hopping processes that must be considered\cite{MF}:
\begin{eqnarray}
\label{Hhop1}
{\cal H}_{hop} = d^\dagger_2 d_1 \biggl[\Delta_0 +
\frac{\Delta_1}{2}
\left(c^\dagger_{1\uparrow}c_{2\uparrow} +
c^\dagger_{1\downarrow}c_{2\downarrow} \right)\biggr. \\ \nonumber
+\Delta_2\biggl.
c^\dagger_{1\uparrow}c_{2\uparrow}c^\dagger_{1\downarrow}c_{2\downarrow}
\biggr] + h.c.
\end{eqnarray}
representing hopping of the (dressed) impurity, jointly with,
respectively, 0, 1 and 2 electrons moving the opposite way. Although we
might start at a high energy scale with negligible $\Delta_1$ and
$\Delta_2$, these will be generated under renormalization and hence
must be included.
In order to analyze the renormalization group flows, it is convenient
to approximate the conduction band $s$-wave electrons $c_{1\epsilon}$
and $c_{2\epsilon}$ by a linear dispersion with a cutoff, at short
times, $\tau_c$, roughly the inverse bandwidth. Then the interactions
with the impurity will essentially be replaced by corresponding phase
shifts, specifically $V_3$ replaced by an effective phase shift that
we denote $\pi Q_0$; $Q_0$ will have the interpretation of an
effective ``charge''. We then have, after
inserting powers of $\tau_c$ to make the couplings dimensionless and
factors of $\pi$ for convenience,
\begin{equation}
\label{calH}
{\cal H}= {\cal H}_0 + {\cal H}_{int} + {\cal H}_{hop}
\end{equation}
with
\begin{equation}
\label{calHo}
{\cal H}_0= \sum_\sigma \int \frac{d\epsilon}{2\pi} \epsilon
\left[c^\dagger_{1\sigma\epsilon}c_{1\sigma\epsilon} +
c^\dagger_{2\sigma\epsilon}c_{2\sigma\epsilon}\right],
\end{equation}
\begin{eqnarray}
\label{calHint}
{\cal H}_{int} &=& \pi Q_0 \left(d^\dagger_2d_2 -d^\dagger_1d_1\right)
\sum_\sigma \left(c^\dagger_{2\sigma}c_{2\sigma}
-c^\dagger_{1\sigma}c_{1\sigma}\right) \\ \nonumber
&+& \pi y \sum_\sigma
\left(c^\dagger_{1\sigma}c_{2\sigma}
+c^\dagger_{2\sigma}c_{1\sigma}\right)
\end{eqnarray}
and
\begin{eqnarray}
\label{calHhop}
{\cal H}_{hop} = d^\dagger_2 d_1 \biggl[\frac{\Delta_0}{2\pi\tau_c}+
\frac{\Delta_1}{2}
\sum_\sigma c^\dagger_{1\sigma}c_{2\sigma} \biggr.\\ \nonumber
+\biggl. \Delta_2 2\pi\tau_c
c^\dagger_{1\uparrow}c_{2\uparrow}c^\dagger_{1\downarrow}c_{2\downarrow}
\biggr] + h.c.
\end{eqnarray}
We have rescaled the electronic term $V_2$ to a coefficient $y$ which
will play the role of a ``fugacity'' for electronic hops.
\subsection{Symmetries}
It is important at this stage to examine the symmetries of
Eq(\ref{calH}). In addition to time reversal, conservation of the
electrons $\left(c\rightarrow e^{i\phi} c\right)$, conservation of the
impurity $\left(d\rightarrow e^{i\phi} d\right)$, and $SU(2)$ spin
symmetry, the only other symmetry is interchange of the two sites and
the corresponding electronic states ($1\leftrightarrow 2$). Note,
however, that if the only hopping term had been $\Delta_1$, and if $y$
vanished, there would be an {\em extra} artificial symmetry $d_1
\rightarrow e^{i\phi} d_1$, $c_1
\rightarrow e^{i\phi} c_1$, $d_2\rightarrow d_2$, $c_2 \rightarrow
c_2$ corresponding to conservation of $N_1=d^\dagger_1 d_1 +
n_{c_1}$ and similarly $N_2$ {\em separately}, where $n_{c_1}$
is the number of the ``one'' electrons which, in the
absence of the channel mixing term $y$, are independent of the ``two''
electrons. As shown in reference \onlinecite{MF}, even if the
electronic states had
been optimally chosen so that there was no mixing of ``one'' and
``two'' electrons at the Fermi energy, the energy dependence of
scattering off the impurity would generate extra mixing terms in
${\cal H}_{int}$, that cannot simply be expressed in terms of $c_1$ and
$c_2$. These would break the artificial symmetry and under
renormalization generate a $y\left(c^\dagger_1 c_2 + h.c.\right)$
mixing term even in the absence of impurity motion. Thus it is best to
include $y$ from the beginning. (The neglected energy dependent
scattering terms will then not play an important role).
In order to understand the difficulties of reaching the 2CK-like fixed
point, this step is {\em crucial}.
The artificial symmetry in the absence of the channel mixing and
$\Delta_0$, $\Delta_2$ terms, corresponds to a conserved pseudo-spin
$N_1-N_2$ which is the sum of the ``$z$-components'' of the impurity
pseudo-spin $d^\dagger_2d_2-d^\dagger_1d_1$ and an electronic
pseudo-spin $n_{c_2} - n_{c_1}$. This pseudo-spin can
play the role of spin for the two-channel Kondo effect and, indeed,
under renormalization the system will flow to this intermediate
coupling 2CK fixed point if $Q>0$.\cite{Footnote1}
Unfortunately, there is no natural small parameter which keeps the
pseudo-spin symmetry breaking terms small.\cite{Footnote2}
\subsection{Bosonization}
In order to carry out the renormalization group analysis for small
bare impurity hopping rates, it is useful, as is standard, to bosonize
the electronic degrees of freedom, treating the electronic states
$c_{1,2 \epsilon}$ as those of a one-dimensional system with two sets
of right moving electrons with ``wavevectors'' $v_F (k-k_F)\propto
\epsilon$. It is simplest to set the Fermi velocity, $v_F=1$, and
treat $\epsilon$ like a
wavevector index, defining
\begin{equation}
\label{Psij(x)}
c_{j\sigma}\left(x\right) \equiv \int \frac{d\epsilon}{2\pi}
e^{i\epsilon x} c_{j\sigma\epsilon}=\frac{1}{\sqrt{2\pi\tau_c}}
e^{i\Phi_{j\sigma}\left(x\right)}
\end{equation}
so that
\begin{equation}
\label{dPhidx}
c^\dagger_{j\sigma}\left(x\right) c_{j\sigma}\left(x\right)=
\frac{1}{2\pi} \frac{\partial \Phi_{j\sigma}\left(x\right)}{\partial x}
\end{equation}
with $j=1,2$ $\sigma=\uparrow,\downarrow$ and $\Phi_{j\sigma}$ being
the corresponding bosonic degrees of freedom, where we have followed
Emery and Kivelson's notation.\cite{EK} Only $c_{j\sigma}\equiv
c_{j\sigma}\left(x=0\right)$ couples to the impurity. Note that in the
standard expression Eq(\ref{dPhidx}) the left hand side is normal
ordered and therefore the (infinite) uniform charge density does not
appear. Also corrections that vanish as $\tau_c\rightarrow 0$ are
neglected; we will be careful to include the effects of extra terms
when they play an important role.
Since we will later need to be careful to have the proper
anticommutation relations, we must insert extra factors of the form
$\exp\left(i\pi N_\mu\right)$, with $N_\mu\equiv \int dx
\Psi^\dagger_\mu\left(x\right) \Psi_\mu\left(x\right)$, into some of
the bosonized expressions to ensure anticommutations of the different
Fermi fields. These will not play a role as long as no spin-flip
processes occur, and, for the time being we ignore them; the needed
modifications are spelled out in Appendix B.
It is useful to define even and odd components of the Bose fields
\begin{equation}
\label{defPhieo}
\Phi_{e,o \sigma}= \frac{1}{\sqrt{2}}\left(\Phi_{2\sigma} \pm
\Phi_{1\sigma}\right),
\end{equation}
the Hamiltonian then becomes
\begin{eqnarray}
\label{defbosonizedH}
{\cal H} &=& {\cal H}_0 + \frac{Q_0}{\sqrt{2}} \sigma_z \sum_\sigma
\frac{\partial\Phi_{o\sigma}}{\partial x}
+y\sum_\sigma \cos \Phi_{o\sigma} \\ \nonumber
&+&\frac{\Delta_0}{2\pi\tau_c} \sigma_x
+ \frac{\Delta_1}{4\pi\tau_c}
\sum_\sigma \left(\sigma_+
\exp\left[i\sqrt{2}\Phi_{o\sigma}\right] +
h.c.\right) \\ \nonumber
&+&\frac{\Delta_2}{2\pi\tau_c} \left(\sigma_+
\exp\left[i\sqrt{2}\left(\Phi_{o\uparrow}+\Phi_{o\downarrow}\right)
\right] + h.c.\right)
\end{eqnarray}
where in all the coupling terms the bosonic fields are evaluated at
$x=0$ and we use the impurity
pseudo-spin operators
\begin{equation}
\label{defsigmaz}
\sigma_z=d^\dagger_2d_2-d^\dagger_1d_1
\end{equation}
and
\begin{equation}
\label{defsigma+-}
\sigma_+=d^\dagger_2d_1\; ,\; \sigma_-=d^\dagger_1d_2.
\end{equation}
The conduction electron part of the Hamiltonian can be written in
terms of boson creation and annihilation operators
$\phi_\mu\left(\epsilon\right)$ with canonical commutation relation
\begin{equation}
\label{phiphi+comrelation}
\left[ \phi_\mu\left(\epsilon\right),
\phi^\dagger_\nu\left(\epsilon'\right) \right] = 2\pi
\delta_{\mu\nu} \delta\left(\epsilon-\epsilon'\right)
\end{equation}
via
\begin{equation}
\label{Phimuxasafnofphi}
\Phi_\mu\left(x\right)=\int_0^\infty
\frac{d\epsilon}{\sqrt{2\pi\epsilon}} \left[
\phi_\mu\left(\epsilon\right) e^{i\epsilon
x}+\phi^\dagger_\mu\left(\epsilon\right) e^{-i\epsilon x}\right]
e^{-\frac{\epsilon\tau_c}{2}}
\end{equation}
and
\begin{equation}
\label{Hoasafnofphi}
{\cal H}_0 = \sum_\mu \int_o^\infty \frac{d\epsilon}{2\pi} \epsilon
\phi^\dagger_\mu\left(\epsilon\right)\phi_\mu\left(\epsilon\right)
e^{-\epsilon\tau_c}
\end{equation}
which involves positive energy parts only. Here $\tau_c^{-1}$ is the
energy cutoff and $\mu$ represents the various Bose fields, i.e. for
Eq(\ref{defbosonizedH}), $\mu=\left(e\uparrow, e\downarrow, o\uparrow,
o\downarrow\right)$. We see that Eq(\ref{defbosonizedH}) does not
include any terms with
$\Phi_{e\downarrow}$ or $\Phi_{e\uparrow}$. Thus, up to operators that
are irrelevant for weak coupling, the impurity is decoupled from the
even boson fields.
It is convenient, following Emery and Kivelson\cite{EK}, to decompose
the odd field $\Phi_o$ which couples to the impurity, into a spin and
charge part, by
\begin{equation}
\label{Phiosigma}
\Phi_{o\sigma}=\frac{1}{\sqrt{2}} \left(\Phi_c + \sigma \Phi_s\right)
\end{equation}
with $\sigma=\pm$ for spin $\uparrow$, $\downarrow$, respectively. The
second term in Eq(\ref{defbosonizedH}) becomes simply
$Q_0 \sigma_z \frac{\partial\Phi_c}{\partial x}\left(0\right)$. This
term, which represents the difference in phase shifts for electron
scattering off the two positions of the impurity, can be shifted away
by a unitary transformation which changes the naive weak coupling scaling of
the hopping terms; this is the conventional approach
used\cite{Murumatsu,Georges,MF,EK} to derive the weak
coupling flows discussed in the next section.
Although the even fields will not play much role for the time being,
for later purposes we also introduce even fields $\Phi_{ec}$,
$\Phi_{es}$ by
\begin{equation}
\label{Phiec,s}
\Phi_{e\sigma}=\frac{1}{\sqrt{2}} \left(\Phi_{ec} + \sigma \Phi_{es}
\right).
\end{equation}
Note that any sum or difference of any two of the $\Phi_{j\sigma}$,
i.e. those that appear from operators bilinear in electron operators,
can be written as a sum or difference of two of the fields $\Phi_s$,
$\Phi_c$, $\Phi_{es}$, $\Phi_{ec}$ with coefficients of {\em unity};
this enables the method of Emery and Kivelson\cite{EK} to work.
\section{Weak hopping analysis}
\label{sec:Analysisoftheweakcouplingpoint}
In order to connect the various amplitudes at the relatively high energy
scale of the effective Hamiltonian
(Eq(\ref{defbosonizedH})) to their renormalized values at low energies
we must
analyze the weak coupling renormalization group (RG) flow
equations for the amplitudes in Eq(\ref{defbosonizedH}). The
magnitudes of the various terms at the crossover scale to intermediate
coupling will determine which regions of the initial parameter space
can flow near to the 2CK fixed point.
Following the procedure in reference \onlinecite{MF} we transform ${\cal H}$
to $U{\cal H}U^\dagger$ using the unitary operator
\begin{equation}
\label{UtransQo}
U=\exp\left[-i\sigma_z Q_0 \Phi_c\right]
\end{equation}
Subsequently we follow the RG approach described there and obtain
the following flow equations for the various amplitudes, where for
later convenience we introduce
\begin{equation}
\label{defq}
q=\frac{1}{2} -Q_0,
\end{equation}
which lies in the range $(-1/2,1/2)$:
\begin{eqnarray}
\label{weakrgeqns}
\frac{d\Delta_0}{dl}&=&\left(\frac{1}{2}+2q-2q^2\right)\Delta_0 +
y\Delta_1 +O\left(\Delta^3\right) \\ \nonumber
\frac{d\Delta_2}{dl}&=&\left(\frac{1}{2}-2q-2q^2\right)\Delta_2 +
y\Delta_1 +O\left(\Delta^3\right) \\ \nonumber
\frac{d\Delta_1}{dl}&=&\left(\frac{1}{2} -2q^2\right)\Delta_1 +
2y\left(\Delta_0+\Delta_2\right) +O\left(\Delta^3\right) \\
\nonumber
\frac{dq}{dl}&=&-2q\left(\Delta_0^2 +\Delta_2^2-\frac{1}{2} \Delta_1^2
\right)+ \left(\Delta_0^2 -\Delta_2^2\right) \\ \nonumber
\frac{dy}{dl}&=&\Delta_1\left(\Delta_0+\Delta_2\right)
\end{eqnarray}
The important cross-terms in the first three equations in
Eq(\ref{weakrgeqns}) that are proportional to the electronic mixing
term $y$, have a simple physical interpretation: they represent the
effects of an impurity and an electronic hop both occuring within a
short time interval so that, at lower energies, this appears as simply
the corresponding combined process. Note that we have not included
$O\left(y^2\right)$ terms in the above equations; the definition of
these will depend on the RG procedure, and they will not qualitatively
change the behavior. Thus, in the spirit of focusing on the important
processes and terms, we ignore them\cite{Footnote12A}.
Noting that $q$ and $y$ are constant to order
$O\left(\Delta^2\right)$, to analyze the flow for weak hopping, we can
safely set them to their initial values, $q_0$ and $y_0$. Then the
first three equations can be diagonalized exactly; the details are
discussed in Appendix A. The RG eigenvalues for the
hopping terms, about the zero hopping fixed line are
\begin{eqnarray}
\label{rgivalues}
\lambda_\pm&=&\frac{1}{2} -2q^2_0 \pm 2\sqrt{q_0^2+y_0^2} \\ \nonumber
\lambda_0&=&\frac{1}{2} -2q^2_0.
\end{eqnarray}
We now note that for $0\leq Q_0\leq 1$, corresponding to
$\left|q_0\right|\leq \frac{1}{2}$, at least two eigenvalues are
positive so that impurity hopping is always relevant (leading to the
conclusion of {\em absence} of impurity localization\cite{MF});
likewise for other ranges of $Q_0$, there will always be at least two
relevant hopping processes if there is only $s$-wave scattering off
the impurity.
The Kondo temperature, $T_K$, is the energy scale at which the first
of the
impurity hopping processes becomes of order unity-- i.e. of order the
renormalized bandwidth.
The system considered by Vlad\'{a}r and
Zawadowski\cite{Zawadowski1} and Vlad\'{a}r {\em et
al}\cite{Zimanyi1}, essentially
amounts to neglecting $\Delta_2$ and $y$; for small
$\left|q\right|$, which will turn out to be the most interesting case,
this misses part of the physics. The reason is simply that they
neglect one relevant operator $\Delta_2$ which mixes with the other
two to give the correct eigenvalues
(Eq\ref{rgivalues}). Furthermore, as will be seen later,
non-vanishing values of $\Delta_2$ and $y$ are crucial to give the
correct renormalization flows close to the intermediate coupling fixed
point.
If we only kept $Q$ and $\Delta_1$ non-zero, their weak coupling flows
would be (up to
coefficients) like those for $J_z$ and $J_\perp$ for the conventional
Kondo problem. The Kondo scale is then simply
\begin{equation}
\label{tkondo}
\frac{T_K}{W} \sim \left(\frac{\Delta_1}{W}\right)^\frac{1}{\lambda_0},
\end{equation}
after reinserting factors of the bandwidth, $W$.
(For the special value $\Delta_1=2Q$, $T_K\sim e^{-\frac{1}{\Delta_1}}$
like in the well known anti-ferromagnetic Heisenberg Kondo
problem). For this artificial case, at scales below $T_K$ the novel two
channel Kondo physics will indeed appear, as we will discuss later. This
results from the approach, at low energies, of the system to the
intermediate coupling 2CK fixed point.
Unfortunately, the breaking of the artificial pseudo-spin symmetry
leads to the appearance of terms which generically drive the flow {\em
away} from the 2CK fixed point. In order to analyze whether the
system can get near to the 2CK fixed point --- the prerequisite for
observation of non-Fermi liquid behavior---, we must be able to identify
the operators near the 2CK fixed point in terms of the original terms
in the
Hamiltonian; the magnitude of the operators, in particular the relevant
ones, can then be
determined, roughly, by ``matching'' the coefficients at the
crossover scale,
$T_K$, between the weak and intermediate coupling regimes.
{}From Appendix A, we see that the crossover temperature, $T_K$, will
generally be a complicated function of the original parameters. Before
examining the magnitude of the various important terms at $T_K$, we
turn to the
behavior near the 2CK fixed point; this will tell us which terms need
to be small at scales of order $T_K$ for 2CK behavior to obtain.
\section{Intermediate coupling analysis and two channel Kondo fixed point}
\label{sec:Intermediatecouplingfixedpoint}
{}From the weak coupling flow equations in Eq(\ref{weakrgeqns}), it is
apparent that, in the absence of electronic mixing ($y$=0) there is a
special value of the effective impurity charge $Q$: for
$Q=\frac{1}{2}$, corresponding to $q=0$, $\Delta_0$, $\Delta_1$ and
$\Delta_2$ all scale in the same way with eigenvalue
$\lambda=\lambda_0=\lambda_\pm=\frac{1}{2}$. By analogy to the
Toulouse limit of the conventional one channel Kondo problem, it is
thus natural to look for a solvable point that corresponds to $y=q=0$,
inspired by the observation that free Fermi fields have scaling
dimension of $\frac{1}{2}$ and thus might be used to represent all of
the hopping terms that appear for $q=0$. This has been carried out
recently by Emery and Kivelson\cite{EK} who ``refermionize'' the
bosonized operators that appear in the Hamiltonian enabling the
computation of the scaling of the various important operators.
By examining the weak coupling flows, it is apparent that
special behavior might occur when $\Delta_0+\Delta_2=0$.
Defining
\begin{equation}
\label{defDelta+}
\Delta_+\equiv \Delta_0+\Delta_2,
\end{equation}
we see that, at least to the order included in Eq(\ref{weakrgeqns}),
for $y=0$ and $\Delta_+=0$, $q$ flows to zero, $y$ is {\em not}
generated and one might hope that the flow would be towards the 2CK
fixed point. Indeed, the intermediate coupling analysis shows that
this can occur, even if
\begin{equation}
\label{defDelta-}
\Delta_- \equiv
\Delta_0-\Delta_2
\end{equation}
is non-zero so that the artificial pseudo-spin symmetry is broken.
Physically the role of $\Delta_+$ is very surprising. The processes
represented by $\Delta_0$ and $\Delta_2$ are very different and the
definitions which make them dimensionless are clearly cutoff
dependent. Thus the special critical behavior must not in general
occur exactly
at $\Delta_+=0$, since the location --- but not the existence --- of
the critical manifold will be affected by irrelevant operators. In
particular, from the weak coupling flows we can see that a non-zero
$q$ combined with $\Delta_- \neq 0$ {\em will} generate $\Delta_+$,
thus a ``bare'' $\Delta_+$ that is non-zero will be needed for the
flow to go to the 2CK fixed point asymptotically.
In order to understand the intermediate coupling behavior and the
special role of $Q=\frac{1}{2}$, following Emery and Kivelson\cite{EK}
and the analogous Toulouse limit\cite{Toulouse} of the conventional
Kondo problem, we perform a unitary transformation with
\begin{equation}
\label{Utrans1/2}
U=\exp\left[-\frac{i}{2}\sigma_z\Phi_c\right]
\end{equation}
which transforms the Hamiltonian to $\tilde{\cal H}=U{\cal H}U^+$
\begin{eqnarray}
\label{Hbosrotated}
\tilde{\cal H} = {\cal H}_0 &+&
\frac{\Delta_1}{2\pi\tau_c}\sigma_x\cos\Phi_s +
\frac{\Delta_-}{2\pi\tau_c}\sigma_y\sin\Phi_c \\ \nonumber
&+&\frac{\Delta_+}{2\pi\tau_c}\sigma_x\cos\Phi_c
+ 2y\cos\Phi_c\cos\Phi_s -
q\sigma_z \frac{\partial\Phi_c}{\partial x} \\ \nonumber
&+&\frac{u}{\pi}\frac{\partial\Phi_{ec}}{\partial x}
-\frac{w}{2\pi}\sigma_x\cos\Phi_c\frac{\partial\Phi_{ec}}{\partial x}
\end{eqnarray}
where we have combined the terms
\begin{eqnarray}
\label{explanatoryforHbosrotated}
\frac{\Delta_2}{2\pi\tau_c}\left(\sigma_+ e^{i\Phi_c} + \sigma_-
e^{-i\Phi_c}\right) &+&
\frac{\Delta_0}{2\pi\tau_c}\left(\sigma_+ e^{-i\Phi_c} + \sigma_-
e^{i\Phi_c}\right) \nonumber \\
= \frac{\Delta_-}{2\pi\tau_c}\sigma_y\sin\Phi_c &+&
\frac{\Delta_+}{2\pi\tau_c}\sigma_x\cos\Phi_c,
\end{eqnarray}
and abbreviated $\Phi_\mu\left(x=0\right)$ simply by
$\Phi_\mu$. We have also reintroduced some of the coupling terms to
the even fields, in particular the marginal term
$u\frac{\partial\Phi_{ec}}{\partial x}$ that arises from the
impurity--position independent part of the electron--impurity
scattering ($V_1$ in Eq(\ref{Ufinal})) and the term
$w\sigma_x\cos\Phi_c\frac{\partial\Phi_{ec}}{\partial x}$, which arises
from the combination of $u$ and impurity hopping terms $\Delta_+$ and
is irrelevant for weak hopping, will play roles in our analysis. An
additional irrelevant term, $\sigma_y \sin\Phi_s
\frac{\partial\Phi_{es}}{\partial x}$ couples the impurity to the electronic
{\em spin} degrees of freedom, but does not feed back to the other
operators and thus we ignore it for the time being. Its effect will be
discussed further in Appendix C.
\subsection{Symmetries}
Since we expect that the symmetries will play an important role, we
should examine what the original symmetries correspond to in
$\tilde{\cal H}$ and ensure that there are no extra symmetries which
might have arisen from the discarding of operators which were naively
irrelevant, since, as shown in reference \onlinecite{MF},
such procedures, especially
when combined with ``large'' transformations, such as
Eq(\ref{Utrans1/2}), can be dangerous.
\subsubsection{Gauge invariance and spin conservation}
Since there are no spin-flip processes, separate gauge transformations
can be made for each spin
$\Phi_{j\sigma}\left(x\right)\rightarrow\Phi_{j\sigma}\left(x\right) +
\theta_\sigma$, corresponding to $\Phi_{e\sigma}\rightarrow\Phi_{e\sigma} +
\sqrt{2} \theta_\sigma$. This does not play much role as the
$\Phi_{ec}$-field enters only as a derivative
$\frac{\partial\Phi_{ec}}{\partial x}$ and the
$\Phi_{es}$-field decouples from the impurity at the level at
which we work (there is feedback, under renormalization from other
operators involving $\Phi_e$ but these only modify pre-existing terms
in $\tilde{\cal H}$).
However, if the $z$-component of electron spin is {\em not} conserved,
then only the symmetry
$\Phi_{ec}\rightarrow\Phi_{ec} + \theta_{ec}$ remains.
\subsubsection{Periodicity}
The definition of the Fermi fields, Eq(\ref{Psij(x)}), implies that
shifting any $\Phi_{j\sigma}$ by $2\pi$ should leave $\tilde{\cal H}$
unchanged. Depending on whether one or both spin components are
shifted, this implies that (ignoring shifts in $\Phi_{ec}$)
\begin{equation}
\label{Phicchange}
\Phi_c\rightarrow\Phi_c +2\pi
\end{equation}
and
\begin{equation}
\label{Phischange}
\Phi_s\rightarrow\Phi_s +2\pi
\end{equation}
are independent symmetries, as is the combination
\begin{eqnarray}
\label{combination1}
\Phi_c\rightarrow\Phi_c +\pi \;&,&\; \Phi_s\rightarrow\Phi_s +\pi\\
\nonumber
\sigma_x\rightarrow-\sigma_x \;&,&\; \sigma_y\rightarrow-\sigma_y ;
\end{eqnarray}
with the necessity for the simultaneous transformation of
$\sigma_{x,y}$ resulting from the unitary transformation, U, which
involves $\Phi_c$ so that $\Phi_c\rightarrow\Phi_c +\pi$ introduces an
extra \mbox{$\exp\left(-\frac{\pi i}{2}\sigma_z\right)=-i\sigma_z$} factor
into $U$ yielding $\sigma_{x,y}\rightarrow -\sigma_{x,y}$ in
$\tilde{\cal H}$.
\subsubsection{Interchange}
Interchanging sites one and two is equivalent to
\begin{eqnarray}
\label{combination2}
\Phi_c\rightarrow-\Phi_c \;&,&\; \Phi_s\rightarrow-\Phi_s, \\
\nonumber
\sigma_y\rightarrow-\sigma_y \;&,&\; \sigma_z\rightarrow-\sigma_z
\end{eqnarray}
with $\Phi_{ec,es}$ unchanged.
\subsubsection{Spin reversal}
Flipping electron spins is simply $\Phi_s\rightarrow-\Phi_s$ and
$\Phi_{es}\rightarrow -\Phi_{es}$.
\subsubsection{Time reversal}
Time reversal transformations change ingoing to outgoing waves,
thereby yielding
$x\rightarrow-x$, $i\rightarrow -i$, all \mbox{$\Phi_\mu\left(x\right)
\rightarrow-\Phi_\mu\left(-x\right)$} and
$\sigma_y\rightarrow-\sigma_y$. Note that here we are {\em not} time
reversing the spins.
\subsubsection{Artificial extra symmetries}
We now see that, indeed, $\tilde{\cal H}$ in Eq(\ref{Hbosrotated})
with all coefficients non-zero, does {\em not} have any artificial
symmetries.
But as seen earlier, an artificial extra pseudo-spin symmetry is possible:
pseudo-spin conservation mod 2 corresponds to $\Phi_c\rightarrow\Phi_c
+\pi$, and more generally full pseudo-spin symmetry corresponds to
independence of the ``one'' and ``two'' electrons,
i.e. \mbox{$\Phi_c\rightarrow \Phi_c +\theta_c$} with any
$\theta_c$. The terms $w$,
$\Delta_+$, $\Delta_-$ and $y$ all violate this, and it can readily be
seen in the representation of $\tilde{\cal H}$ of
Eq(\ref{Hbosrotated}) that $y$ combined with
$\Delta_1$ generates $\Delta_+$, and $q$ combined with $\Delta_-$
generates $\Delta_+$, as expected.
In the representation of Eq(\ref{Hbosrotated}) we see that there is
another possible artificial symmetry: If $w$, $\Delta_+$, $y$ and $q$ are
all zero, then
\begin{equation}
\label{Phictrans}
\Phi_c\rightarrow\pi-\Phi_c
\end{equation}
becomes a symmetry. This, as we shall see, restricts the system
automatically to the stable critical manifold of the 2CK fixed point.
But note that because of the unitary transformation of
Eq(\ref{Utrans1/2}) $\Phi_c\rightarrow\pi-\Phi_c$ does {\em not}
correspond to a realizable symmetry in terms of the original variables
since it mixes hops involving the impurity alone and those involving
the impurity together with two electrons. (Indeed many other
irrelevant terms neglected in $\tilde{\cal H}$ will also violate this
artificial symmetry.)
\subsection{Refermionization}
If $\Delta_+$, $q$, $y$ and the other operators neglected in
$\tilde{\cal H}$ all vanish, then the system will still exhibit the
novel intermediate coupling 2CK behavior. As we shall see, the only
relevant operator near the 2CK fixed point, which is consistent with
the true symmetries of the impurity hopping between two equivalent
sites, is $\Delta_+$, thus only one combination of physical quantities
needs to be adjusted to obtain the 2CK behavior. Unfortunately, this
is a combination which is not naturally small.
The behavior near the 2CK fixed point can most easily be found
following Emery and Kivelson\cite{EK} by ``refermionizing'' the Bose
fields $\Phi_c$, $\Phi_s$ and $\Phi_{ec}$ that appear in $\tilde{\cal
H}$ in Eq(\ref{Hbosrotated}), noting that $\exp i\Phi_\mu$, (with
$\Phi_\mu$ properly normalized) is like some pseudo-Fermi field. The
details are discussed in Appendix B.
Crudely, for $\mu=c$, $s$, $ec$, $es$, each field $e^{i\Phi_\mu}$ is
replaced by a new Fermi field, $\Psi_\mu$, and $\sigma_-$ by a local
Fermi field $d$, with appropriate factors of $e^{i\pi N_\mu}$ and
$e^{i\pi N_d}$ to give the correct anticommutation relations. The
symmetries can most easily be seen, and the Hamiltonian simplified, by
writing the new Fermi fields in terms of a set of Majorana (hermitian)
fermions:
\begin{eqnarray}
\label{defMajoranafermions}
d&=&\frac{1}{\sqrt{2}} \left(\gamma+i\delta\right) \\ \nonumber
\Psi_\mu\left(x=0\right)&=&\frac{1}{\sqrt{2}}\left(\alpha_\mu+
i\beta_\mu\right).
\end{eqnarray}
Note that
\begin{equation}
\label{defsigmazwrtgammadelta}
\sigma_z=2i\gamma\delta.
\end{equation}
The symmetry restrictions can now be examined in terms of these
variables; the details are given in Appendix B. The periodicity of
the Bose fields $\Phi_c$ and $\Phi_s$ simply implies that only terms
with an even number of Fermi fields can appear in the Hamiltonian.
{\em Gauge invariance} implies that $\Psi_{ec}$ can only appear as
$\Psi_{ec}^\dagger\Psi_{ec}$ and the {\em $z$-component of spin
conservation} that $\Psi_{es}$ cannot appear in the absence of
magnetic fields in the $x$- or $y$-direction. {\em Spin reversal},
because of the role of the ordering operators, takes
$\Psi_s\rightarrow-\Psi_s^\dagger$ and
$\Psi_{es}\rightarrow-\Psi_{es}^\dagger$, implying that $\alpha_s$ and
$\alpha_{es}$ by themselves are excluded by spin reversal symmetry.
{\em Interchange symmetry} takes $\Psi_c\rightarrow - \Psi_c^\dagger$,
$\Psi_s\rightarrow -\Psi_s^\dagger$ and $d\rightarrow d^\dagger$
thereby requiring that $\alpha_c$ and $\delta$ must appear together.
Finally, {\em time reversal} takes $x\rightarrow -x$, $i\rightarrow
-i$ and $\Phi\rightarrow -\Phi$, allowing only real coefficients of
$\Psi_\mu$ operators, and hence forbidding terms like $i\gamma\alpha$.
The Hamiltonian becomes
\begin{eqnarray}
\label{Hrefermionized1}
\hat{\cal H}= {\cal H}_0 &+& \frac{i}{\sqrt{2\pi\tau_c}}
\left(\Delta_1 \gamma \beta_s + \Delta_-
\gamma \beta_c + \Delta_+ \delta\alpha_c\right) \\ \nonumber
&-& 4\pi y \gamma\delta\alpha_c\beta_s + 4\pi q
\gamma\delta\alpha_c\beta_c\\ \nonumber
&+& 2iu\alpha_{ec}\beta_{ec} +
w\sqrt{2\pi\tau_c}\delta\alpha_c\alpha_{ec}\beta_{ec}
\end{eqnarray}
with ${\cal H}_0$ the kinetic energy of the four (eight Majorana) new
Fermi fields, $\Psi_\mu$. With the full symmetries of the system, the
other five fields ($\alpha_s$ and the $ec$-, $es$-fields), cannot
appear in the couplings to the
impurity, except in relatively innocuous forms, involving the simple
potential coupling to the average position of the impurity,
$i\alpha_{ec}\beta_{ec}$ and combinations of this with other terms, as
well as the irrelevant term $\delta\alpha_s\alpha_{es}\beta_{es}$,
which will be discussed in Appendix C.
Note that other potentially important operators, like
$\delta\beta_s\alpha_c\beta_c$ and $i\delta
\frac{\partial\alpha_c\left(0\right)}{\partial x}$ are excluded by
time reversal invariance.
The original 2CK problem studied by Emery and Kivelson\cite{EK} and
Sengupta and Georges\cite{Georges} corresponds to
$\Delta_-=\Delta_+=y=w=0$. In our case non-zero $\Delta_-$ can be
important by observing that ``half'' of the impurity, $\gamma$,
couples to both $\beta_c$ and $\beta_s$, thus it is convenient to
rediagonalize and make linear combinations of these, $\beta_I$ and
$\beta_X$ yielding, with
\begin{equation}
\label{defDeltaK}
\Delta_K=\sqrt{\Delta_1^2 + \Delta_-^2},
\end{equation}
\begin{eqnarray}
\label{Hrefermionized2}
\hat{\cal H}= {\cal H}_0 &+&
\frac{i}{\sqrt{2\pi\tau_c}} \left(\Delta_K
\gamma \beta_I +
\Delta_+ \delta\alpha_c \right)\\ \nonumber
&+& 4\pi\bar{q} \gamma\delta\alpha_c\beta_X +4\pi\bar{y}
\gamma\delta\alpha_c\beta_I \\ \nonumber
&+& 2iu\alpha_{ec}\beta_{ec} +\sqrt{2\pi\tau_c}w
\delta\alpha_c\alpha_{ec}\beta_{ec},
\end{eqnarray}
where $\bar{y}$ and $\bar{q}$ are linear combinations of the original
$y$ and $q$ (see Eq(\ref{defqybar}) in Appendix B for details). The
above rediagonalization of $\beta_c$ and $\beta_s$ roughly
corresponds to a rotation of ``spin'' axes in the conventional 2CK
language.
{}From the electronic kinetic energy ${\cal H}_0$, the $\alpha$'s and
$\beta$'s all scale, with time scale $\tau$, as
$\tau^{-\frac{1}{2}}$. If all the couplings are
small, then the anti-commutation relations of $\gamma$ and $\delta$
imply that they are dimensionless so that $\Delta_+$ and $\Delta_K$
scale as $\tau^{-\frac{1}{2}}$, while $\bar{q}$ and $\bar{y}$ are
marginal as from the weak coupling
analysis of Section II (Eq(\ref{weakrgeqns})) and $w$ is irrelevant.
\subsection{Two channel Kondo fixed point and flows}
When $\bar{y}=\bar{q}=\Delta_+=w=0$,
the Hamiltonian in Eq(\ref{Hrefermionized2}) corresponds to the Toulouse
limit analyzed by Emery and Kivelson\cite{EK}. As a free fermion
system it can be analyzed straightforwardly. In this limit ``half'' of
the impurity, $\delta$, is uncoupled from the electrons and thus has
no dynamics, while the other ``half'', $\gamma$, gets dynamics from
coupling to the electrons with correlations at large imaginary times
\begin{equation}
\label{gammacorrelation}
\left<T_\tau\gamma\left(\tau\right)\gamma\left(0\right)\right> \sim
\frac{1}{\tau}.
\end{equation}
Together these yield the non-Fermi liquid 2CK behavior
\begin{equation}
\label{sigmazcorrelation}
\left<T_\tau\sigma_z\left(\tau\right)\sigma_z\left(0\right)\right> \sim
\frac{1}{\tau}.
\end{equation}
It is important to note that\cite{NozieresPrivComm} the solvable
Hamiltonian is {\em not} generally at the 2CK fixed point. Indeed, the
correlations are readily seen to exhibit crossover from weak coupling
behavior, $\left<\sigma_z \sigma_z\right>\sim const$, to the non-Fermi
liquid behavior of Eq(\ref{sigmazcorrelation}) for
$\tau\gtrsim\Delta_K^{-2} \tau_c$.
The 2CK fixed point, formally, corresponds to
$\Delta_K\rightarrow\infty$. It is more convenient, however, to allow
instead the normalization of $\gamma$ to change, corresponding to
letting the coefficient of the $\int \gamma\partial_\tau \gamma \:
d\tau$ in the Lagrangian vary. At the fixed point, this coefficient,
say $g_\gamma$, will be zero, while $\Delta_K$ becomes a constant; the
correlations of $\gamma$ are then simply the inverse (in frequency
space) of those of $\beta_I$, i.e. a pure power law Eq(\ref{Ggamma}).
To connect the two regimes together, one could choose to renormalize
so that, for example, $\frac{\Delta_K^2}{4\pi}+ g_\gamma=1$ [a
particularly convenient choice; (see Eq(\ref{DeltaKgconstraint})], by
rescaling $\gamma$ under renormalization by a $\Delta_K$-dependent
amount. Details about this procedure are given in Appendix B. [Note,
however, that the resulting fixed point Hamiltonian (with
$\Delta_K=\sqrt{4\pi}$) will {\em not } have the pseudo-spin $SU(2)$
symmetry of the pure 2CK problem. This is because the RG approach we
have implemented here, including the unitary transformation of
Eq(\ref{Utrans1/2}), is inherently anisotropic in channels (equivalent
to spin of conventional 2CK problem). In Appendix C we will show that
this anisotropy will not affect the results.]
At the 2CK fixed point, the above scaling implies
$\gamma\sim\tau^{-\frac{1}{2}}$, while $\delta$ is still dimensionless
so that the RG eigenvalues of the other operators can be read off
immediately: +1/2 for $\Delta_+$, the unique relevant operator
consistent with the symmetries of the problem; -1/2 for $\bar{q}$,
$\bar{y}$ and $w$, which are the three leading irrelevant operators
discussed so far (the fourth will appear in Appendix C); and 0 for
$u$, which is marginal but redundant, in that it does not affect the
impurity dynamics.
The operator corresponding to $\bar{q}$ does not give rise to any
terms which couple $\delta$ linearly to the $\alpha$'s and $\beta$'s
and hence will not generate $\Delta_+$. It
is the leading irrelevant operator for the 2CK problem identified by
Sengupta and Georges\cite{Georges}. An artificial symmetry is
responsible for its special role, indeed just the one discussed
earlier: Eq(\ref{Phictrans}),
$\Phi_c\rightarrow \pi-\Phi_c$ corresponds to the discrete symmetry
$\alpha_c\rightarrow-\alpha_c$ and $\beta_X\rightarrow-\beta_X$ which
does not have a natural representation even in terms of $\Phi_c$ and
$\Phi_s$. But if present, this artificial symmetry excludes the
generation of terms like $\Delta_+$, $\bar{y}$ and $w$, even in the
presence of $\bar{q}$. In fact, without these terms, the extra
artificial symmetry is really an $O(2)$ symmetry in the
$\{\alpha_c, \beta_X\}$ pair, consisting of a $U(1)$ of
rotations in the $\left(\alpha_c, \beta_X\right)$ plane, combined (in a
non-commutative way) with $Z_2$, the usual site interchange symmetry, which
takes $\alpha_c\rightarrow-\alpha_c$ and
$\beta_X\rightarrow\beta_X$. This is exactly analogous to the $O(2)$
symmetry present in the model treated by
Emery and Kivelson\cite{EK} and
Sengupta and Georges\cite{Georges}.
In contrast, the irrelevant operators $w$ and $\bar{y}$, break the
artificial symmetry and yield, at lowest order, the generation of the
relevant operator $\Delta_+$ (see Eq(\ref{intermediatergeqns}))
\begin{equation}
\label{Delta+rgeq}
\frac{d\Delta_+}{dl}= \frac{1}{2} \Delta_+ + 2\bar{y}\Delta_K +
\frac{wu}{\sqrt{2}\left(1+u^2\right)},
\end{equation}
consistent with expectations from weak coupling. This implies as
stated earlier, that the critical point will not be exactly at
$\Delta_+=0$.
Of the three leading irrelevant operators, it should be noted that,
although $\bar{y}$ has the same scaling dimension as $\bar{q}$ and
$w$, it has a different role close to the 2CK fixed point. The reason
is that it couples to the term $i\gamma\beta_I$, already present at
the fixed point. But the $\Delta_K$ term at the fixed point suppresses
fluctuations of $\beta_I$, causing the leading term in the
correlations of $\beta_I$ to {\em vanish} at the fixed point (with
sub-dominant terms caused by $g_\gamma\neq 0$). As a result, unlike
$\bar{q}$ and $w$ which each yield $O\left(T\ln T\right)$
contributions to the impurity specific heat\cite{Georges} --- a key
feature of a 2CK non-Fermi liquid --- the singular part arising from
$\bar{y}$ is only of $O\left(T^3\ln T\right)$, for temperatures $T\ll
T_K$. Thus only {\em two} of the above independent leading irrelevant
operators give leading singular specific heat corrections. In fact,
there is a third one, involving only spin degrees of freedom, which is
discussed in Appendix C.
In Section VI, the conditions for accessibility of the 2CK fixed point
are analyzed. We turn here to further analysis of the behavior near
the 2CK fixed point.
\subsection{Symmetry Breaking Operators}
Up to this point we have only dealt with an electron-impurity system
which is invariant under $Z_2$ ($1\leftrightarrow 2$ interchange) and
spin $SU(2)$ symmetry. However, in a realistic situation of an
impurity in a metal, there will generally be a non-zero, although
possibly small asymmetry between the two sites which will break the
$Z_2$ symmetry. Furthermore, in the presence of a magnetic field, the
equivalence between the two spin channels will be lost, leading to a
situation similar to the anisotropic Kondo
problem\cite{Nozieres1,Nozieres2,Coleman}. As might be expected, in
both cases, the symmetry breaking terms are relevant and in their
presence the system flows away from the 2CK fixed point. In this
section, we will briefly comment on the effects of symmetry breaking
terms close to the 2CK fixed point.
It is clear from the discussion in the previous section that for an
operator to be relevant close to the 2CK fixed point, it has to be of
the form $i\delta \chi$ where $\chi$ is a Majorana fermion of scaling
dimension 1/2. From the ten Majorana fermions (four pairs of
$\alpha_\mu$, $\beta_\mu$ and $\gamma$, $\delta$, all listed in Table
\ref{Table1}) we can make nine such operators. Excluding $i\delta
\alpha_{ec}$ and $i\delta \beta_{ec}$ due to total electron number
conservation (which only allows $\alpha_{ec}$ and $\beta_{ec}$ to
appear together as $\alpha_{ec}\beta_{ec}$) we are left with seven
possible terms.
From the transformation properties under the discrete symmetries of
the system, listed in Table \ref{Table1}, it can be seen that
$\beta_c$ and $\beta_s$ have the same symmetries; indeed this is why
the $\Delta_1$ and $\Delta_-$ terms in Eq(\ref{Hrefermionized1}) could
be combined into the $\Delta_K$ term of Eq(\ref{Hrefermionized2}). In
the presence of small $i\delta\beta_c$ and $i\delta\beta_s$ terms, a
small rotation of the ($\delta$, $\gamma$) pair as well as a small
additional rotation of the ($\beta_c$, $\beta_s$) pair can be
performed to yield just a slightly modified $i\Delta_K\gamma\beta_I$
term, and a single remaining relevant perturbation, the $i\delta\beta_X$
term. The extra operators ($i\gamma\beta_X$ and $i\delta\beta_I$) are
thus ``redundant''.\cite{Footnote22A} Therefore at the 2CK fixed
point, there are exactly {\em six relevant} operators, all with RG
eigenvalue of 1/2, like $\Delta_+$. These, along with their symmetry
properties, are listed in Table \ref{Table2}.
The first three operators in Table \ref{Table2} do not break the
spin $SU(2)$ symmetry. Among these, the first, our familiar
$i\Delta_+\delta\alpha_c$, is interchange and time-reversal symmetric,
corresponding to the relevant part of the channel pseudo-spin operator
$S_x$. Correspondingly, the second, $i\delta\gamma$, breaks
interchange symmetry but is time reversal invariant, corresponding to a
$S_z$ operator. This will result from simple asymmetry between the
impurity energies of the two sites, i.e. a $\sigma_z$ term in the
original Hamiltonian. Finally, $i\delta\beta_c$ (or, equivalently,
$i\delta\beta_s$) breaks interchange and time reversal which makes it an
imaginary operator, i.e. it is generated by a $S_y$ operator in the
channel sector, corresponding to complex hopping matrix elements.
The relevant spin $SU(2)$ breaking operators are $i\delta\alpha_s$,
$i\delta\alpha_{es}$ and $i\delta\beta_{es}$; these correspond to
joint electron-impurity hops accompanied by a spin flip or carrying
electronic spin. They correspond to combinations of
$\sigma_+\left(c^\dagger_{1\uparrow}c_{2\uparrow} -
c^\dagger_{1\downarrow}c_{2\downarrow} \right)$, $\sigma_+
c^\dagger_{1\downarrow}c_{2\uparrow}$,
$\sigma_+c^\dagger_{1\uparrow}c_{2\downarrow}$ and their hermitian
conjugates and are discussed in Appendix B (see Eq(\ref{Hsf}) and
Eq(\ref{MajoranaHsf})). The first corresponds to the ``flavor''
anisotropy term in the conventional two channel Kondo
model\cite{Nozieres1,Nozieres2,Coleman}, and, being interchange and
time reversal symmetric but odd under spin flip, is induced by a
magnetic field in the $z$-direction. Interestingly, the remaining two
relevant spin $SU(2)$ breaking operators are {\em odd} under the $Z_2$
interchange transformation. This means that in order to adjust them to
zero in a non-zero external magnetic field, one would have to tune
also terms that break the interchange symmetry of the problem, making
any additional novel, finite magnetic field, non--Fermi liquid fixed
points (analogous to that found in zero field for $\Delta_+\gg
\Delta_1$,$\Delta_-$), extremely hard to observe in a system that does
not have interchange symmetry.
\section{Additional intermediate coupling fixed point}
\label{Additionalfixedpoint}
In the previous section we analyzed the behavior of the system close
to the 2CK fixed point, that corresponds to the limit $\Delta_K \gg
\Delta_+$. There, we showed that $\gamma$, ``half'' of the impurity,
acquired non-trivial dynamics, which essentially gave rise to the
non--Fermi liquid behavior of the system. However, as is evident by
examining Eq(\ref{Hrefermionized2}), it should be, in principle,
possible to get the same type of non--Fermi behavior if the inequality
above were reversed ($\Delta_K \ll \Delta_+$).
Indeed, following the same arguments analyzed above, we see that when
$\Delta_K=\bar{q}=\bar{y}=w=u=0$ and $\Delta_+ \neq 0$ we have a
critical point, which has an extra artificial $U(1)$ symmetry, namely
rotations in the $\left(\beta_I, \beta_X\right)$ plane. At this
secondary fixed point, $\delta$, the other ``half'' of the impurity,
acquires dynamics, rather than $\gamma$. The operators $\bar{q}$,
$\bar{y}$, $i\gamma\beta_I\Psi_{ec}^\dagger\Psi_{ec}$ and
$i\gamma\beta_X\Psi_{ec}^\dagger\Psi_{ec}$ have scaling dimension 3/2
and thus are irrelevant with RG eigenvalue -1/2. But of these, only
the last two will give singular specific heat corrections
($O\left(T\ln T\right)$).
However, there is an important difference from the primary 2CK fixed
point discussed earlier. About this fixed point, there are {\em two}
relevant operators, consistent with the symmetries of the model,
namely $i\gamma \beta_I$ and $i\gamma \beta_X$ (or, equivalently,
$i\gamma \beta_c$ and $i\gamma \beta_s$), both with dimension 1/2.
These correspond simply to $\Delta_1$ and $\Delta_-$ before the change
of variables (Eq(\ref{defbetaix})) leading to
Eq(\ref{Hrefermionized2}); they can be generated from nonzero $q$ and
$y$ . The existence of two relevant operators is due to the fact that
$\beta_I$ and $\beta_X$ transform the {\em same} way under the
discrete symmetries; thus the artificial $U(1)$ symmetry {\em cannot}
be extended into an $O(2)$ group (as was the case for the 2CK fixed
point). As a result, this new fixed point is harder to find in the
interchange symmetric case than the primary 2CK fixed point, as it
requires the impurity-single-electron hopping term to be small and the
two-electron- plus-impurity and the simple impurity hopping terms to
be almost exactly equal at the Kondo scale. It should be noted,
however, that the {\em total} number of relevant, dimension 1/2,
operators around this fixed point is again six, including the above
mentioned ones, together with $i\gamma\delta$ and the three spin
$SU(2)$ symmetry breaking operators, $i\gamma\alpha_s$,
$i\gamma\alpha_{es}$ and $i\gamma\beta_{es}$, as discussed in Section
IV.E and listed in Table \ref{Table2}, indicating that the fixed point
symmetry is again that of the conventional 2CK model. This is
supported by the fact that, just like in the case of the 2CK fixed
point, there are four operators with scaling dimension 3/2, albeit
with completely different symmetries.
Finally, some comments are needed on the the nature of this novel
fixed point. We should first stress the {\em absence} of $\Delta_1$,
the impurity-single-electron hopping term, which together with the
$Q_0$ term (see Eq(\ref{calHint})) would form the conventional
Kondo-like interaction term. Hence, the appearance of non-Fermi liquid
behavior does {\em not} originate from the competition between the
spin up and spin down electrons to form a channel-pseudo-spin singlet
ground state with the impurity, but, rather, in the presence of strong
impurity-electron repulsion ($Q_0\approx 1/2$), from the competition
between bare impurity tunnelling and two-electron-plus-impurity
tunnelling. Formally, there is an analogy with the conventional two
channel flavor-anisotropic Kondo model,\cite{Nozieres2} with
$\Delta_\pm$ playing the role of the Kondo couplings of the two
``flavor'' channels, which can be seen in the left hand side of
Eq(\ref{explanatoryforHbosrotated}) if $\Phi_c$ is substituted by
$\Phi_s$. When $\Delta_0\neq\Delta_2$ one flavor channel couples more
strongly to the impurity therefore screening it alone at low energy,
which results in usual Fermi liquid behavior. However, if
$\Delta_0=\Delta_2$ the flavor anisotropy disappears and the system
flows to a non-Fermi liquid fixed point (provided $\Delta_1$ is zero).
As a result, although this fixed point may be in the same universality
class as the conventional 2CK model, the mechanism that brings it
about is completely different physically.
\section{Accessibility of the two channel Kondo fixed point and Conclusions}
In the previous sections we have shown how the physical operators in
the tunnelling impurity problem behave near the two channel Kondo
fixed point. In particular, we observed that a linear combination,
$\Delta_+=\Delta_0+\Delta_2$, of the bare impurity hopping,
$\Delta_0$, and the impurity-plus-two-electron hopping, $\Delta_2$, is
relevant and drives the system away from this special critical point
resulting in conventional Fermi liquid behavior at low temperatures,
as for the usual one-channel Kondo system. If one could somehow tune
$\Delta_+$, (or one other coupling such as the electronic hopping $y$)
then one might be able to tune through the critical point and find the
2CK non-Fermi liquid behavior at low temperatures in the vicinity of a
critical coupling. Unfortunately, such tuning over an adequate range
is probably difficult to achieve. Thus one probably has to rely on the
hope that a natural regime of couplings will lead to flow under
renormalization close to the 2CK fixed point. Vlad\'{a}r and
Zawadowski\cite{Zawadowski1} appear to suggest that this should be the
case. Unfortunately, more complete analysis implies the converse, that
only for fortuitous reasons would the impurity system --- even without
asymmetry between the sites --- exhibit 2CK behavior at low $T$.
In this last section, we use the weak coupling analysis of Section III
and Appendix A combined with the intermediate coupling analysis of
Section IV and Appendix B, to find criteria for approaching close to
the 2CK fixed point.
Since we are interested in systems in which the Kondo temperature is
much less than the bandwidth, the weak hopping behavior will control
the relative strengths of couplings at the Kondo scale, at which the
first of the hopping terms becomes of order the renormalized bandwidth.
Operators which are irrelevant for weak hopping will flow away rapidly
under renormalization, changing by finite amounts the remaining
parameters, $q$, $y$ and the $\left\{\Delta_i\right\}$, $i=0$, 1, 2.
For example, the complicated hopping-scattering term, $w$, which was
discussed in Section IV and plays a role near the intermediate
coupling fixed point, will be of order the hopping terms
$\left\{\Delta_i\right\}$ or smaller initially and flow away rapidly,
modifying, among other terms, $\Delta_+$ as in Eq(\ref{Delta+rgeq}),
in the process. This will be the main role of such a term and we can
incorporate its effects into a modified ``bare'' $\Delta_0$ and
$\Delta_2$. We thus start at an energy scale substantially below the
bandwidth at which the important parameters for $Q\in\left[0,1\right]$
are just $q$, $y$ and the $\left\{\Delta_i\right\}$ at this scale, the
irrelevant operators having become small. The relevant eigenvalues
for the hopping about the zero hopping fixed manifold will be
universal. For small $y$ they are given by Eq(\ref{rgivalues}).
If $Q$ is initially small, i.e. $q\leq 1/2$, $\lambda_+$ will be
substantially larger than the other eigenvalues and thus a particular
linear combination of the $\left\{\Delta_i\right\}$ will grow fast.
Unfortunately, as can be seen from Appendix A, this combination
($\tilde{\Delta}_0$) is the wrong one to yield flow near the 2CK fixed
point as it {\em includes} $\Delta_+$. Only if this combination,
$\tilde{\Delta}_0$, is initially very small relative to a power of
another linear combination of the same $\left\{\Delta_i\right\}$, can
behavior near the 2CK critical point be obtained; furthermore the
criteria become more stringent the lower the Kondo temperature, as
shown in Appendix A.
Better prospects occur when $Q\approx1/2$ (i.e. $q$
small). Unfortunately, even if the one-electron plus impurity hopping
term, $\Delta_1$, were initially much bigger than $\Delta_0$ and
$\Delta_2$, the purely electronic hopping term $y$ --- determined
basically by the spatial separation of the two impurity sites\cite{MF}
--- would combine with $\Delta_1$ to generate the {\em wrong}
combination, $\Delta_+$, of $\Delta_0$ and $\Delta_2$, as in
Eq(\ref{Delta+rgeq}). Thus, again, unless $y$ is small the criteria
from Appendix A are very strict, as could be anticipated from the $y$
dependence of $\lambda_\pm$.
The best prospects are thus for $q$ and $y$ both small so that the
eigenvalues are all comparable. But this is just the condition for the
analysis of Section IV and Appendix B via refermionization to be
valid. To get near the 2CK fixed point, $\Delta_+$ must remain small,
thus we can study the RG equations Eq(\ref{intermediatergeqns}) to
leading order in $\Delta_+$ and the linear combinations of $y$ and
$q$, $\bar{y}$ and $\bar{q}$;
\begin{eqnarray}
\label{intermedrgeqnstext2}
\frac{d\Delta_K}{dl} &=&
\frac{\Delta_K}{2}\left(1-\frac{\Delta_K^2}{4\pi}\right)
\\ \nonumber
\frac{d\bar{q}}{dl} &=& -\frac{\Delta_K^2}{8\pi}\bar{q}
\\ \nonumber
\frac{d\bar{y}}{dl} &=& -\frac{\Delta_K^2}{8\pi}\bar{y}
\end{eqnarray}
and, ignoring $w$ from Eq(\ref{Delta+rgeq})
\begin{equation}
\label{Delta+intermediatergeq}
\frac{d\Delta_+}{dl} = \frac{1}{2} \Delta_+ + 2\bar{y}\Delta_K.
\end{equation}
At the intermediate coupling 2CK fixed point, the convention we have
chosen yields $\Delta_K^*=\sqrt{4\pi}$, so that $\bar{y}$ and
$\bar{q}$ have the correct eigenvalues there as well as for weak
coupling ($\Delta_K\approx 0$).
Integrating the above equations we
find that the criterion to flow to the 2CK fixed point is, to leading
order in $\Delta_+$ and $\bar{y}$
\begin{equation}
\label{critsurfacecriterionw/outw}
\Delta_+ -
\frac{2\bar{y}\Delta_K}{1-\frac{\Delta_K^2}{4\pi}}
\ln\left|\frac{\Delta_K^2}{4\pi}\right| =0
\end{equation}
The parameters can be evaluated at any scale where
the intermediate coupling RG equations are valid, i.e. for small
$\Delta_+$, $\bar{y}$ and $\bar{q}$. If these are small at the
starting scale, then their starting values can be used. Note that
since $\Delta_K=\sqrt{\Delta_1^2+\left(\Delta_0-\Delta_2\right)^2}$,
unless $\Delta_0$ and $\Delta_2$ are almost exactly equal and
opposite, $\Delta_K$ will be at least as big as $\Delta_+$ initially
so that $\left|\bar{y}\ln\Delta_K\right|$ needs to be small rather than
just $y$, even in the best case of $q=1/2-Q$ small (recall that
$\bar{y}$ is a linear combination of $q$ and $y$ given by
Eq(\ref{defqybar})).
We thus see that the flows near the 2CK
fixed point, in particular the generation via Eq(\ref{Delta+rgeq}) of
the unique relevant operator, $\Delta_+$ from other operators, yield
stringent conditions for the accessibility of the non-Fermi liquid 2CK
behavior. If there is no symmetry between the two sites, then the
presence of a second relevant operator (see Section IV) makes
prospects even worse.
This work strongly suggests that to observe two-channel-Kondo-like
non-Fermi liquid behavior of an impurity hopping between two sites in a
metal one must either be able to tune some parameters over a
substantial range, or be extremely lucky. This casts doubt on the
interpretation of Ralph {\em et al}\cite{Ralph1,Ralph2} of their narrow
constriction tunnelling data. One possibility, although perhaps
farfetched, is that these might be some kind of defects tunnelling in
environments with higher symmetry, or at least approximate
symmetry. In another paper, we will show how an impurity hopping among
{\em three} sites with triangular symmetry can, without fine tuning,
lead to a two channel Kondo behavior at low temperatures.
\acknowledgments
We would like to thank Andreas Ludwig, Jinwu Ye, Dan
Ralph, Jan von Delft, Igor Smolyarenko and especially Anirvan Sengupta
for useful discussions. This work was supported by the National
Science Foundation via grant DMR 91-06237.
\end{multicols}
|
\section{Introduction}
Many physical objects exhibit some form of symmetry. Most galaxies for
instance, have axes or planes of symmetry. The motivation for this study is
that a symmetric equilibrium configuration generally is the outcome of the
evolution from an asymmetric state. We would like to trace the effect of the
asymmetries.
A problem is that studies of the evolution of actual physical systems are
difficult and so relatively rare. We propose therefore to ignore, at least for
the time being, the actual physical mechanisms and to consider systems
described by a Hamiltonian of the form
\begin{equation}
\mathcal{H}(p, q, \epsilon t) = \mathcal{H}_s(p, q) + a(\epsilon t)
\mathcal{H}_a(p, q)
\end{equation}
where $\mathcal{H}_s$ is the part of the Hamiltonian which is symmetric in some
sense; $\mathcal{H}_a$ is the asymmetric
part which is slowly vanishing as we put
\begin{equation}
a(0) = 1, \lim_{t \rightarrow \infty} a(\epsilon t) = 0, 0 < \epsilon \ll 1
\end{equation}
To study the dynamics induced by the Hamiltonian $\mathcal{H}(p, q, \epsilon
t)$ is still a formidable problem. So we
simplify as much as possible to obtain
\begin{equation}
\label{Sys:x}
\left\{
\begin{array}{rcl}
\dot{x}_1 & = & x_2 \\
\dot{x}_2 & = & -x_1 + a(\epsilon t) x_1^2
\end{array}
\right.
\end{equation}
which is derived from the one degree of freedom Hamiltonian
\begin{equation}
\mathcal{H}(p, q, \epsilon t) = \frac{1}{2} (p^2 + q^2) + \frac{1}{3}
a(\epsilon t) q^3
\end{equation}
identifying $p = x_2$, $q = x_1$. We shall associate with system (\ref{Sys:x})
the ``unperturbed'' system which
arises for $\epsilon = 0$
\begin{equation}
\label{Sys:x:eps=0}
\left\{
\begin{array}{rcl}
\dot{x}_1 & = & x_2 \\
\dot{x}_2 & = & -x_1 + x_1^2
\end{array}
\right.
\end{equation}
\begin{figure}
\begin{center}
\includegraphics[height=8cm]{unperturbed.eps}
\end{center}
\caption{The dynamics of the unperturbed system (\ref{Sys:x:eps=0})}
\label{Fig:unperturbed system}
\end{figure}
We note that in the autonomous system (\ref{Sys:x:eps=0}) there are basically
two regions (figure \ref{Fig:unperturbed system}): within the homoclinic
solution the orbits are bounded, outside the homoclinic solution the orbits
diverge to infinity (with the exception of the stable manifold and the saddle
point itself). In system (\ref{Sys:x}) we have no fixed saddle point, still it
turns out that we have two
separate regions of initial values in which the orbits are bounded or diverge
to infinity.
Since the dynamics of systems (\ref{Sys:x}) and (\ref{Sys:x:eps=0}) are the
same on an $O(1)$ timescale, it is instructive (though slightly wrong) to view
system (\ref{Sys:x}) as having a saddle point moving slowly towards infinity
and having a slowly expanding homoclinic orbit. Within this picture, an orbit
can remain bounded in two ways, either by starting inside the homoclinic orbit,
or by getting ``captured'' by the slowly expanding homoclinic orbit, which can
only happen if the orbit starts sufficiently close to the stable manifold of
the saddle point.
Using a special transformation we shall discuss the boundaries of these regions
in section \ref{Section: The boundary of the stable part of phase space}.
A special case, $a(\epsilon t) = \exp(-\epsilon t)$ can be studied easily and
help us to understand the general case.
In section \ref{Section: Averaging inside the stable region} we perform
averaging in the so-called stable region where bounded solutions are found.
This involves the use of elliptic and hypergeometric functions, rather hard
analysis, where we are
supported by Mathematica 2.2 running under SunOS 4.1.3. \\
After determining the validity of the averaged equation we establish the
existence of an adiabatic invariant in the stable region, valid for all time.
Even more remarkable is that explicit calculations of this invariant show that
the evolution of
phase points will show significant traces of its asymmetrical past for all
time.
In section \ref{Section: The boundary layer} we need subtle reasoning to
discuss what is going on in the boundary layer near
the boundary of the stable domain.
The analysis in this paper is based on averaging methods but, because of its
direct relation to dissipative mechanics
(section \ref{Section: The boundary of the stable part of phase space}), it
clearly profits from the results by
Haberman and Ho~\cite{H&H 1,H&H 2} and Bourland and Haberman~\cite{B&H}. At the
same time our analysis should be placed within the theory of adiabatic
invariants, which has been summarized recently in an admirable survey by
Henrard~\cite{H}.
We finally note that in the context of galactic dynamics, some rather different
examples based on classical results of the theory of adiabatic invariants were
given by Binney and Tremaine~\cite{B&T}.
\section{The boundary of the stable part of phase space}
\label{Section: The boundary of the stable part of phase space}
As we explained in the introduction, the phase space of system (\ref{Sys:x})
can be separated into two parts. Since we are dealing with a time-dependent
system, we must specify the time for which a particular separation holds. We
use the following definition:
The \emph{stable} part of phase space consists of the points $(x_1, x_2)$, for
which the orbit $\gamma (x_1, x_2, 0)$ starting in $(x_1, x_2)$ at $t=0$
remains bounded for $t$ going to infinity. All other points define the
\emph{unstable} part of phase space.
Clearly, a point $(x_1, x_2)$ can only be contained in the stable region if it
lies within an $O(\epsilon )$ neighbourhood of the area bounded by the
homoclinic orbit of system (\ref{Sys:x:eps=0}). If this is not the case,
$\gamma (x_1, x_2, 0)$ will reach the upper branch of the unstable manifold of
the saddle point of system (\ref{Sys:x:eps=0}) in a finite time and clear off
to infinity.
We must not overlook the orbits starting close to the lower branch of the
stable manifold of the saddle point of system (\ref{Sys:x:eps=0}), which can
reach the just described $O(\epsilon )$ neighbourhood too. It will turn out
that although this region may look small, it produces the major part of the
stable region.
These considerations help us locating the boundary of the stable region
approximately.
The location of the boundary of the stable region separates the part of phase
space in which all orbits diverge to infinity ($(x_1, x_2) \rightarrow{}
(+\infty , +\infty )$) from the part of phase space in which the orbits tend to
circle around the origin for $t$ going to infinity, so if we expect to see
effects of the vanishing of the asymmetric potential somewhere, it is just
within this boundary.
The key step in analyzing system (\ref{Sys:x}) is performing the transformation
\begin{equation}
\label{Trafo:x->y}
y_1 = a(\epsilon t) x_1
\end{equation}
The idea behind this transformation is to try to fix the ``slowly moving saddle
point'' of system (\ref{Sys:x}).
Demanding that $\dot{y}_1 = y_2$, we arrive at the system
\begin{equation}
\label{Sys:y}
\left\{
\begin{array}{rcl}
\dot{y}_1 & = & y_2 \\
\dot{y}_2 & = & -y_1 + y_1^2 + 2 \epsilon \frac{a'(\epsilon
t)}{a(\epsilon t)} y_2 + {\epsilon}^2 \left( \frac{a''(\epsilon
t)}{a(\epsilon t)} - 2 \frac{{a'(\epsilon t)}^2}{{a(\epsilon t)}^2}
\right) y_1
\end{array}
\right.
\end{equation}
where $a'(\epsilon t)$ stands for ${\left. \frac{\mathrm{d}a(\xi
)}{\mathrm{d}\xi } \right\vert}_{\xi{} = \epsilon t}$ and similarly for
$a''(\epsilon t)$.
By transformation (\ref{Trafo:x->y}) the slow time-dependence has moved to
$O(\epsilon)$ terms; still, system (\ref{Sys:y}) looks more complicated than
system (\ref{Sys:x}).
However, we will be able to neglect the $O({\epsilon}^2)$ term in most of our
calculations.
We should also note that system (\ref{Sys:y}) is \emph{not} Hamiltonian
anymore, since we have applied a non-canonical transformation. Indeed the
$O(\epsilon )$ term is a friction term, causing the origin $(y_1, y_2) = (0,
0)$ to become an attracting focus instead of a center.
In the analysis of system (\ref{Sys:y}) we start with a special choice of
$a(\epsilon t)$.
\subsection{The special case $a(\epsilon t) = e^{-\epsilon t}$}
We will first calculate the location of the boundary of the stable region for
the special, but physically important case
\begin{equation}
\label{Special:e-macht}
a(\epsilon t) = e^{-\epsilon t}
\end{equation}
We will show later that the general case does not differ much from this special
case.
With the choice (\ref{Special:e-macht}) for $a(\epsilon t)$, system
(\ref{Sys:y}) reduces to
\begin{equation}
\label{Sys:y:e-macht}
\left\{
\begin{array}{rcl}
\dot{y}_1 & = & y_2 \\
\dot{y}_2 & = & -y_1 + y_1^2 - 2 \epsilon y_2 - {\epsilon}^2 y_1
\end{array}
\right.
\end{equation}
It is remarkable that for this special yet interesting choice of $a(\epsilon
t)$, our system becomes autonomous, which reduces the calculations because the
dependence on the initial time has vanished into the transformation
(\ref{Trafo:x->y}). \\
We also note that we have succeeded in fixing the saddle point: The saddle
point of system (\ref{Sys:y:e-macht}) is located in $(1 + {\epsilon}^2, 0)$.
The saddle point not being located in $(1,0)$ as we intended would introduce a
lot of extra small terms in our calculations. To avoid these we map the saddle
point onto $(1,0)$ by substituting $y_i \rightarrow{} (1 + {\epsilon}^2) y_i,
i= 1, 2$, to obtain
\begin{equation}
\label{Sys:y:e-macht:mapped}
\left\{
\begin{array}{rcl}
\dot{y}_1 & = & y_2 \\
\dot{y}_2 & = & -(1 + {\epsilon}^2) (y_1 - y_1^2) - 2 \epsilon y_2
\end{array}
\right.
\end{equation}
So we have reduced the calculation of the boundary of the stable region of
system (\ref{Sys:x}) to the calculation of the (time-independent) region of
attraction of system (\ref{Sys:y:e-macht:mapped}).
It is easily seen (figure \ref{Fig:y:e-macht:manif}) that the region of
attraction of system (\ref{Sys:y:e-macht:mapped}) is bounded by the stable
manifold of the saddle point.
\begin{figure}
\begin{center}
\includegraphics[height=8cm]{fig1.eps}
\end{center}
\caption{The stable and unstable manifold of system
(\ref{Sys:y:e-macht:mapped}) with $\epsilon = 0.1$}
\label{Fig:y:e-macht:manif}
\end{figure}
It is well known that generally the stable manifold of a perturbed system (with
parameter $\epsilon$) lies in an $O(\epsilon )$ neighbourhood of the stable
manifold of the unperturbed system (with $\epsilon = 0$).
The unperturbed system
\begin{equation}
\label{Sys:y:e-macht:mapped:eps=0}
\left\{
\begin{array}{rcl}
\dot{y}_1 & = & y_2 \\
\dot{y}_2 & = & -y_1 + y_1^2
\end{array}
\right.
\end{equation}
is simple and totally understood. It has a first integral
$E(\epsilon = 0)$ where
\begin{equation}
\label{Def:integral:eps=0}
E(\epsilon = 0) = \frac{1}{2} y_2^2 + \frac{1}{2} y_1^2 - \frac{1}{3} y_1^3
\end{equation}
and the unstable manifold coincides with the homoclinic orbit $E(\epsilon = 0)
= \frac{1}{6}$.
Using $E(\epsilon = 0)$ in our calculations for the perturbed system would
introduce some higher order terms. Instead, we extend the
definition of $E$ with suitable $O(\epsilon^2)$ terms which cancel these
terms. Again, this is only for calculational convenience.
\begin{equation}
\label{Def:integral:eps}
E = \frac{1}{2} y_2^2 + \frac{1}{2} (1 + {\epsilon}^2) y_1^2 - \frac{1}{3} (1
+ {\epsilon}^2) y_1^3
\end{equation}
It is instructive to combine figure \ref{Fig:y:e-macht:manif} with the
homoclinic orbit of the unperturbed system (\ref{Sys:y:e-macht:mapped:eps=0}),
which produces figure \ref{Fig:y:e-macht:manif+homocl}.
\begin{figure}
\begin{center}
\includegraphics[height=8cm]{fig2.eps}
\end{center}
\caption{The homoclinic orbit (represented by the thin line) of system
(\ref{Sys:y:e-macht:mapped:eps=0}) added to figure \ref{Fig:y:e-macht:manif}}
\label{Fig:y:e-macht:manif+homocl}
\end{figure}
We will now approximate the location of the stable manifold of the saddle point
of system (\ref{Sys:y:e-macht:mapped}) by calculating the variation of $E$
along the stable manifold. Since this variation is an $O(\epsilon )$ effect, we
may use the unperturbed stable manifold in this calculation, which involves
elliptic functions. From this variation of $E$ along the stable manifold, we
can deduce the location of the perturbed stable manifold.
If we follow the flow along the stable manifold from a point $(y_{10}, y_{20})$
to a point $(y_{11}, y_{21})$ we get:
\begin{equation}
\frac{\mathrm{d}E}{\mathrm{d}t} = -2 \epsilon y_2^2
\Rightarrow
\Delta E = \int{-2 \epsilon y_2^2 dt} = -2 \epsilon
\int_{y_{10}}^{y_{11}}{y_2 dy_1}
\end{equation}
The integral appearing in this expression has to be calculated with
$O(\epsilon)$ precision, which allows us to substitute the explicitly known
orbits of the unperturbed system (\ref{Sys:y:e-macht:mapped:eps=0}).
These orbits ($y_2(y_1)$) are readily obtained from the definition of the
first integral (\ref{Def:integral:eps=0}).
To calculate the variation of $E$ along the upper branch of the stable
manifold,
we take $(y_{11}, y_{21})$ = $(1,0)$ and get, after some analysis as indicated
above:
\begin{equation}
E(y_1) = \frac{1}{6} (1 + {\epsilon}^2) + 2 \epsilon \left( \frac{3}{5} +
\frac{\sqrt{3} (y_1 - 2) {(2 y_1 + 1)}^{(3/2)}}{15} \right) + O({\epsilon}^2
{(y_1 - 1)}^2)
\end{equation}
which is valid for $-\frac{1}{2} \leq y_1 \leq 1$ and $y_2 > 0$.
For $y_1 < -\frac{1}{2}$ we have $\dot{E} = O({\epsilon}^2)$ and therefore we
get:
\begin{equation}
E(y_1) = \frac{1}{6} + \frac{6}{5} \epsilon + O({\epsilon}^2)
\end{equation}
which is valid for $y_1 < -\frac{1}{2}$
Taking $(y_{11}, y_{21})$ = $(-\frac{1}{2}, y_{21})$ we get:
\begin{equation}
\label{Eq:log:epsilon:error}
E(y_1) = \frac{1}{6} + 2 \epsilon \left( \frac{3}{5} - \frac{\sqrt{3} (y_1 -
2) {(2 y_1 + 1)}^{(3/2)}}{15} \right) + O({\epsilon}^2 \log \epsilon)
\end{equation}
which is valid for $-\frac{1}{2} \leq y_1 \leq 1$ and $y_2 < 0$.
The special form of the error term arises from the fact that the homoclinic
orbit is only an $O(\sqrt{\epsilon})$ approximation of the stable manifold
for $y_1$ close to 1 and negative $y_2$ (just under the saddle point).
This follows from the analysis in Haberman and Ho~\cite{H&H 1}.
Taking $(y_{11}, y_{21})$ = $(1, y_{21})$ we get
\begin{equation}
E(y_1) = \frac{1}{6} + 2 \epsilon \left( \frac{9}{5} + \frac{\sqrt{3} (y_1 -
2) {(2 y_1 + 1)}^{(3/2)}}{15} \right) + O({\epsilon}^2 \log \epsilon)
\end{equation}
which is valid for $y_1 > 1$ and $y_2 < 0$.
To calculate the variation of $E$ along the lower branch of the stable
manifold,
we take $(y_{11}, y_{21})$ = $(1,0)$ and making use of the expressions for the
explicitly known lower branch of the stable manifold of the unperturbed system
(\ref{Sys:y:e-macht:mapped:eps=0}) we find
\begin{equation}
E(y_1) = \frac{1}{6} (1 + {\epsilon}^2) + 2 \epsilon \left( \frac{3}{5} +
\frac{\sqrt{3} (y_1 - 2) {(2 y_1 + 1)}^{(3/2)}}{15} \right) + O({\epsilon}^2
{(y_1 - 1)}^2)
\end{equation}
which is valid for $y_1 > 1$ and $y_2 < 0$.
So, we have now calculated the variation of $E$ all over the stable manifold of
the saddle point of system (\ref{Sys:y:e-macht:mapped}). What is left to do is
to deduce the location of the stable manifold itself from this variation, which
is not very hard.
Given a value of $y_1$, one first calculates the corresponding value of $E$
using the appropriate formula given above. Using the definition of $E$
(\ref{Def:integral:eps}), one calculates the corresponding value of $y_2$. This
amounts to solving a third order polynomial, which can even be done
explicitly.
In particular one can compute the intersection of the stable manifold with the
$y_1$-axis, which occurs (approximately) in $(-\frac{1}{2} - \frac{8}{5}
\epsilon , 0)$.
\subsection{The boundary of the stable region for arbitrary $a(\epsilon t)$}
We now return to the discussion of the general system (\ref{Sys:y}).
It turns out that the analysis is essentially the same as for the
special case $a(\epsilon t) = e^{-\epsilon t}$
We claim that the behaviour of system (\ref{Sys:y}) is (with a certain error)
described by the system
\begin{equation}
\label{Sys:y:equiv}
\left\{
\begin{aligned}
\dot{y}_1 &= y_2 \\
\dot{y}_2 &= -y_1 + y_1^2 + 2 \epsilon \frac{a'(0)}{a(0)} y_2
\end{aligned}
\right.
\end{equation}
\emph{as far as the location of the boundary of the stable region is
concerned}.
The idea behind this statement is that if an orbit of system (\ref{Sys:y})
starts at a distance
$O(\delta )$ inside the \emph{unstable} region, it will reach the upper branch
of the unstable manifold after an interval of time $O(\log \delta )$, because
it has to pass
the saddle point at a distance $O(\delta )$ (sometimes twice). \\
Using Gronwall's inequality, it is easy to show that the orbit of system
(\ref{Sys:y:equiv}) starting at the same initial point, will diverge at most
$O(\frac{\epsilon^2 \log \delta}{\delta})$ from the exact orbit. \\
Since we know the boundary of the stable region with precision
$O({\epsilon}^2 \log \epsilon )$, we must take $\delta$ to be larger than
$O({\epsilon}^2 \log \epsilon )$ for our calculations to make sense. \\
Consequently, the orbit of system (\ref{Sys:y:equiv}) will diverge at most
$o(1)$ from the exact orbit and will thus diverge to infinity too. \\
Thus, a starting point $(y_1, y_2)$ lying more than $O({\epsilon}^2 \log
\epsilon )$
inside the unstable region produces an orbit diverging to infinity both in
system
(\ref{Sys:y}) and in system (\ref{Sys:y:equiv}).
We can apply exactly the same argument to the stable region, which proves that
the boundary of the stable region of system (\ref{Sys:y}) coincides with the
boundary of the stable region of system (\ref{Sys:y:equiv}) up to
$O({\epsilon}^2 \log \epsilon )$.
So we have the important conclusion that, to calculate the boundary of the
stable region of system (\ref{Sys:y}),
we can use the formulas derived for the special case
$a(\epsilon t) = e^{-\epsilon t}$ with $\epsilon$ replaced by
$-\frac{a'(0)}{a(0)} \epsilon$.
\section{Averaging inside the stable region for arbitrary $a(\epsilon t)$}
\label{Section: Averaging inside the stable region}
Knowing the location of the boundary of the stable region we proceed to study
the stable region itself (the unstable region is clearly not very
interesting). We have to do this study in two parts in which we consider the
interesting dynamics which takes place close to (we will make this more
precise) the boundary of the stable region (i.e. in the \emph{boundary layer})
and in the inner domain. At a safe distance from the
boundary layer, system (\ref{Sys:y}) will
behave more and more like a harmonic oscillator. The natural way to approach
such a problem is to apply the theory of averaging.
\subsection{Averaging in the inner domain}
Averaging in the vicinity of the origin is a simple exercise involving
averaging over harmonic functions. This is not what we have in mind; we shall
average over a part of the inner domain as large as possible. This involves
averaging over elliptic functions.
\subsubsection{Calculation of the averaged equation}
To perform averaging, we need one or more quantities with a small
($O(\epsilon)$) time derivative, i.e. which depend slowly on time. A
natural candidate for this quantity is the exact integral
(\ref{Def:integral:eps=0}) of the unperturbed system, for which we have
\begin{equation}
\label{timeder:integral:eps=0}
\frac{\mathrm{d}E}{\mathrm{d}t} = 2 \epsilon \frac{a'(\epsilon t)}{a(\epsilon
t)} y_2^2 +
\epsilon^2 \left( 1 - y_1 + \left( \frac{a''(\epsilon t)}{a(\epsilon t)} - 2
\frac{{a'(\epsilon t)}^2}{{a(\epsilon t)}^2} \right)
\right)
y_1 y_2
\end{equation}
To be able to average this equation, we have to put restrictions on $a(\epsilon
t)$:
\begin{equation}
\label{restr:average:a}
\begin{aligned}
\frac{a'(\xi)}{a(\xi)} &\textrm{ is bounded for all positive } \xi \\
\frac{a''(\xi)}{a(\xi)} &\textrm{ is bounded for all positive } \xi
\end{aligned}
\end{equation}
Most decaying functions of interest satisfy these restrictions. Functions
decaying extremely
rapidly, such as $a(\xi) = \exp(-\exp(\xi))$, do not satisfy these
restrictions. But since $a(\xi)$ decays very rapidly, we can safely put
$a(\xi)$ equal to zero for all $\xi$ bigger than some $\xi_0$ for which
$a(\xi_0) \ll 1$, without affecting the dynamics. Other examples of
functions which do not satisfy (\ref{restr:average:a}) are
functions which vanish in a finite time like $a(\xi) = 1 - \xi$.
Again we can restrict the time span such that this poses no problem.
To average equation (\ref{timeder:integral:eps=0}), we consider $\tau =
\epsilon t$ as an independent variable and add the equation
\begin{equation}
\dot{\tau} = \epsilon
\end{equation}
Since we only have to average the $O(\epsilon)$ part of equation
(\ref{timeder:integral:eps=0}), we have to average $y_2^2$ along a periodic
orbit of the unperturbed system (\ref{Sys:y:e-macht:mapped:eps=0}). This
amounts to calculating the integral of $y_2^2$ along the periodic orbit and
involves the period of the periodic orbit. This is in the spirit of
averaging as for instance presented in Sanders en Verhulst~\cite{S&V}.
To calculate $\int{y_2^2}dt$, we make use of $\dot{y}_1 = y_2$, which
reduces the calculation to the action $\int{y_2}dy_1$. The functional
dependence of
$y_2$ on $y_1$ for the unperturbed system can be retrieved from the exact
integral (\ref{Def:integral:eps=0}) and is the square root of a third order
polynomial. \\
Using this, we find that we also need this standard integral
\begin{equation}
\int_{a}^{b}{\sqrt{(x-a)(b-x)(c-x)}} = \frac{1}{24} \sqrt{6} \pi {(b - a)}^2
\sqrt{c - a} \sideset{_2}{_1}F\left(-\frac{1}{2}, \frac{3}{2}, 3, \frac{b -
a}{c - a}\right)
\end{equation}
with $\sideset{_2}{_1}F$ the hypergeometric function, which holds when $a \leq
b \leq c$. \\
The $a$, $b$ and $c$ are the exacts roots of a third order polynomial and
are thus awkward expressions even for our simple unperturbed problem.
Surprisingly, the combinations $b-a$ and $c-a$ reduce to manageable
expressions:
\begin{equation}
\label{b-a:c-a}
\begin{aligned}
b - a &= \sqrt{3} \sin\left(\frac{1}{3} \arcsin(12 E - 1) +
\frac{\pi}{6}\right) \\
c - a &= \sqrt{3} \cos\left(\frac{1}{3} \arcsin(12 E - 1)\right)
\end{aligned}
\end{equation}
Substituting all this we get
\begin{equation}
\label{integral:y2:over:y1}
\begin{aligned}
\int{y_2}dy_1 = &2 \frac{1}{24} \sqrt{6} \pi 3 \sin^2\left(\frac{1}{3}
\arcsin(12 E - 1) + \frac{\pi}{6}\right) \times \\
&\times \sqrt{\sqrt{3} \cos\left(\frac{1}{3} \arcsin(12 E - 1)\right)}
\times \\
&\times \sideset{_2}{_1}F\left(-\frac{1}{2}, \frac{3}{2}, 3,
\frac{\sin\left(\frac{1}{3} \arcsin(12 E - 1) +
\frac{\pi}{6}\right)}{\cos\left(\frac{1}{3} \arcsin(12 E -
1)\right)}\right)
\end{aligned}
\end{equation}
The factor 2 arises because we have to integrate once from $b$ to $a$ and once
from $a$ to
$b$.
To calculate the period of the periodic orbit of the unperturbed system, we
apply the standard technique of separation of variables to the exact integral
(\ref{Def:integral:eps=0}).
This leads us through a calculation similar to the one above, resulting in:
\begin{equation}
\label{period:unperturbed:orbit}
\begin{aligned}
\mathrm{period} = &2 \sqrt{6} \frac{1}{\sqrt{\sqrt{3} \cos\left(\frac{1}{3}
\arcsin(12 E - 1)\right)}} \times \\
&\times \mathrm{K}\left(\frac{\sin\left(\frac{1}{3} \arcsin(12 E - 1) +
\frac{\pi}{6}\right)}{\cos\left(\frac{1}{3} \arcsin(12 E -
1)\right)}\right)
\end{aligned}
\end{equation}
where K is the complete elliptic integral of the first kind.
We finally obtain the averaged equation by dividing equation
(\ref{integral:y2:over:y1}) by
equation (\ref{period:unperturbed:orbit}) and adding some extra factors from
equation (\ref{timeder:integral:eps=0}):
\begin{equation}
\label{averaged:equation}
\begin{aligned}
\dot{\bar{E}} &= 2 \epsilon \frac{a'(\epsilon t)}{a(\epsilon t)}
\frac{\int{y_2}dy_1}{\mathrm{period}} \\
&= \epsilon \frac{a'(\epsilon t)}{a(\epsilon t)} A(\bar{E})
\end{aligned}
\end{equation}
It does not add much to the understanding of the problem to write down the
averaged equation in
it's full form. That is why we omit this.
All that matters is that the right hand side is an explicitly known function
$A(\bar{E})$ of $\bar{E}$,
which we can approximate to arbitrary precision, and of time.
\subsubsection{Validity of the averaged equation}
Since the averaged equation is an approximation of the exact system
(\ref{Sys:y}), we have to
address the question of the accuracy of this approximation, on which timescale
it holds and
where in the stable region.
We expect that the closer we start to the homoclinic orbit of the unperturbed
system, the less accurate the averaged equation will become. The dynamics
splits up in two qualitatively different time intervals, the first in which the
orbit slowly separates from the homoclinic orbit and the second in which the
orbit
slowly spirals towards the attracting origin.
We start with the first time-interval. As we will show in section
\ref{subsection:Approaching the boundary layer}, apart from a sub-boundary
layer of size
$O(\exp(-\frac{1}{\epsilon}))$, this time-interval has a size of
$O(\frac{1}{\epsilon})$ \emph{independent} of the initial distance from the
homoclinic orbit.
The total error introduced by the averaging process in the first time-interval
is of
$O(\epsilon T_0)$ ($T_0$ is the period of the unperturbed orbit corresponding
to $\bar{E}(0)$,
the initial value of $\bar{E}$). A short explanation of this estimate is given
in section \ref{Argument: Averaging breaks down}.
For the second interval we can make use of the standard averaging theorems,
from which we get
that the introduced error on the second interval is of $O(\epsilon)$ and that
we are allowed
to extend the second interval to infinity, because all orbits are attracted to
the origin (see Sanders en Verhulst~\cite{S&V}, chapter 4). \\
This attracting property of the orbits also implies that the error introduced
from the first
does not blow up. Therefore, the total error introduced by the averaging
process is of
$O(\epsilon T_0)$ valid for all time.
As we will also show in section \ref{subsection:Approaching the boundary
layer}, for orbits
starting close to the homoclinic orbit of the unperturbed system $\bar{E}(0) =
\frac{1}{6}$,
we have that $T_0$ is of order $-\log(\frac{1}{6} - \bar{E}(0))$, which implies
that the averaged equation can be used to approximate the dynamics up to a
distance of
$O(\exp(-\frac{1}{\epsilon}))$ from the homoclinic orbit of the unperturbed
system.
More quantitative details on the boundary layer will be given
in section \ref{subsection:Approaching the boundary layer}.
We will also see in section \ref{subsection:Approaching the boundary layer}
that the averaged
equation indeed breaks down when we approach the boundary layer.
\subsubsection{Analysis of the averaged equation}
We now turn to the analysis of the averaged equation
(\ref{averaged:equation}).
The first
thing one should notice is that the effect of the decaying function $a(\epsilon
t)$ can be
removed from the equation by transforming to the new time $\tau$
\begin{equation}
\label{averaged:transf:tau}
\left\{
\begin{aligned}
\tau &= -\frac{1}{\epsilon} \log(a(\epsilon t)) \\
a(\epsilon t) &= e^{-\epsilon \tau}
\end{aligned}
\right.
\end{equation}
Note that this transformation reduces to the identity transformation in the
special case
$a(\epsilon t) = e^{-\epsilon t}$. \\
It is remarkable that, given condition (\ref{restr:average:a}), it is not
important at all how $a(\xi)$ decays to zero, the dynamics of the
system does not change, apart from a rescaling of the time axis.
Applying this transformation produces the autonomous, 1-dimensional system
\begin{equation}
\label{averaged:equation:tau}
\frac{\mathrm{d}\bar{E}}{\mathrm{d}\tau} = -\epsilon A(\bar{E})
\end{equation}
We can solve this system explicitly by separation of variables, but
unfortunately
we do not have a primitive of $\frac{1}{A(\bar{E})}$ in the form of an
elementary function. \\
But we can draw some important conclusions from this system, of which the most
important
one is the existence of an \emph{adiabatic invariant}: As noted, it is
always possible to solve system (\ref{averaged:equation:tau}), which gives
the solution $\bar{E} = \bar{E}(\bar{E}(0), \epsilon \tau)$ as a function of
the initial
condition and slow time. Again, in principle one can solve this equation for
$\bar{E}(0)$
as a function of $\bar{E}$ and $\epsilon \tau$. Inverting the time
transformation
(\ref{averaged:transf:tau}), one finds $\bar{E}(0)$ as a function of $\bar{E}$
and
$\epsilon t$. Since $\bar{E}(0)$ is obviously time-independent, we reach the
conclusion that
\vspace{2ex}
\emph{There exists a global adiabatic invariant inside the homoclinic orbit of
the unperturbed system with the exclusion of an exponentially thin boundary
layer,
valid for all time, determined by equation (\ref{averaged:equation:tau}).}
\vspace{2ex}
For special cases we are able to produce these calculations explicitly, which
we will show now.
To understand these cases well, it is important to know how $\int{y_2}dy_1$,
the period and $A$
depend on $\bar{E}$. This is shown in figure \ref{Fig:dependence:E}.
\begin{figure}
\begin{center}
\includegraphics[height=7.5cm]{fig3.eps}
\end{center}
\caption{The dependence of $\int{y_2}dy_1$, the
$\frac{\mathrm{period}}{2\pi}$ and $A$ on $\bar{E}$}
\label{Fig:dependence:E}
\end{figure}
It is clear that $\int{y_2}dy_1$ depends almost linearly on $\bar{E}$
throughout the entire
interval. This is understandable, since it is similar to the dependence of the
area of a disk on
its radius. What is not transparent is that the derivative of this function
goes to infinity as
$\bar{E}$ goes to $\frac{1}{6}$, but so slowly that its integral remains
bounded. \\
The period is close to $2\pi$ for small $\bar{E}$ as it should be, because in
this region the
unperturbed system behaves nearly like a harmonic oscillator with $\omega =
1$.
When $\bar{E}$ goes to $\frac{1}{6}$, the period goes to infinity, because the
orbits are approaching the saddle point, in the neighbourhood of which they
will stay a long time for
each passage. \\
The quotient of the two, $A$, shows the linear behaviour of $\int{y_2}dy_1$ for
small $\bar{E}$,
because the period is almost constant. However, $A$ has a maximum
($0.248320\ldots$) at $\bar{E} = 0.152640\ldots$, after which
it rapidly drops to zero. We could have predicted that $A$ is small for
$\bar{E}$ close to
$\frac{1}{6}$, since all the time the orbits are close to the saddle point, the
righthand side
of equation (\ref{timeder:integral:eps=0}) is small ($y_2 \ll 1$), resulting in
a small average.
\subsubsection{The adiabatic invariant}
We now turn to the calculation of the adiabatic invariant for $\bar{E}(0)$
small (we will make this
more precise later on). \\
To approximate the adiabatic invariant, we perform a Taylor expansion of
$A(\bar{E})$ around 0.
We note that the hypergeometric function forces us to use $\sqrt{\bar{E}}$
as expansion variable instead of just $\bar{E}$.
However, it turns out that the coefficients in front of the non-integer
powers of $\bar{E}$ are equal to zero, at least to fifth order. After a long
calculation we arrive at the following expansion, valid for $0 \leq \bar{E} <
\frac{1}{6}$
\begin{equation}
A(\bar{E}) = 2 \bar{E} - \frac{5}{6} {\bar{E}}^2 - \frac{155}{54} {\bar{E}}^3
- \frac{61135}{5184} {\bar{E}}^4 - \frac{825409}{15552} {\bar{E}}^5 +
O({\bar{E}}^{5\frac{1}{2}})
\end{equation}
To approximate the adiabatic invariant, we truncate the series after the second
order terms, since
we are interested in the first non-trivial deviation from a slowly attracting
focus.
Substituting this quadratic expression into the averaged equation
(\ref{averaged:equation:tau}) we get
\begin{equation}
\frac{\mathrm{d}\bar{E}}{\mathrm{d}\tau} = -2 \epsilon \bar{E} + \frac{5}{6}
\epsilon {\bar{E}}^2
\end{equation}
which is easy to solve giving
\begin{equation}
\bar{E}(\epsilon \tau) = \frac{2
\bar{E}(0)}{\left(2-\frac{5}{6}\bar{E}(0)\right)e^{2 \epsilon \tau} +
\frac{5}{6}\bar{E}(0)}
\end{equation}
From this we readily obtain the adiabatic invariant
\begin{equation}
\frac{\bar{E}(0)}{2 - \frac{5}{6}\bar{E}(0)} = \frac{\bar{E}(\epsilon
\tau)}{2 - \frac{5}{6}\bar{E}(\epsilon \tau)} e^{2 \epsilon \tau}
\end{equation}
We are now able to specify what we meant with $\bar{E}(0)$ small. Since we have
neglected
$O({(\bar{E}(0))}^3)$ terms, we have introduced a new error of order
${(\bar{E}(0))}^3$ in the approximation of the solution.
Since we do not want this error term to dominate the error
introduced by the averaging process ($O(\epsilon)$), we take $\bar{E}(0)$ to be
$O({\epsilon}^{1/3})$.
Expanding the adiabatic invariant around $\bar{E} = 0$, we see that the first
non-trivial
correction to the slowly attracting focus (with adiabatic invariant $\bar{E}(0)
= \bar{E} e^{2\epsilon\tau}$) is given
by $\frac{5}{48}{\bar{E}}^2 e^{2\epsilon\tau}$ resulting in a slightly slower
collapse onto the
origin $(y_1, y_2) = (0, 0)$. \\
These arguments hold for the $(y_1, y_2)$ phase space only. To extend them to
the
original $(x_1, x_2)$ phase space, we have to invert the time-transformation
(\ref{averaged:transf:tau}) and the phase space transformation
(\ref{Trafo:x->y}), after
which we obtain the adiabatic invariant in the $(x_1, x_2)$ phase space:
\begin{equation}
{{3 a(\epsilon t) {x_1^2} -
2 {{a(\epsilon t)}^2} {x_1^3} +
6 \epsilon a'(\epsilon t) x_1 x_2 +
3 a(\epsilon t) {x_2^2} }\over
{72 a(\epsilon t) - 15 {{a(\epsilon t)}^3} {x_1^2} +
10 {{a(\epsilon t)}^4} {x_1^3} -
30 \epsilon {{a(\epsilon t)}^2} a'(\epsilon t) x_1 x_2 -
15 {{a(\epsilon t)}^3} {x_2^2}}}
\end{equation}
We include this rather lengthy expression, because it reveals an important
phenome\-non: Due to the
cross-terms $x_1 x_2$, the level curves of the adiabatic invariant
\emph{for a fixed time} ``resemble'' ellipses, of which the long axis and the
short axis differ by an $O(\epsilon)$ amount. and which are rotated around the
origin, causing asymmetry with respect to the $y_1$ and $y_2$ axis. \\
We did expect this for finite time, but this behaviour persists when we let $t$
tend to
infinity. Put in other words, when $t$ goes to infinity, our dynamical system
(\ref{Sys:x}) becomes symmetric (with respect to $x_1$ and $x_2$), but the
level curves of the adiabatic
invariant remain asymmetric. We have reached this important conclusion:
\vspace{2ex}
\emph{The evolution of an ensemble of phase points towards a symmetric
potential will show significant (i.e. $O(\epsilon)$) traces
of its asymmetrical past, for all time.}
\vspace{2ex}
So there is a sort of hysteresis effect present: although the
system becomes symmetric, it still ``knows'' that it was asymmetric in the
past. \\
We note that this phenomenon is not present in the $(y_1, y_2)$ phase space,
where the level curves of the adiabatic invariant are symmetric with respect to
the $y_1$-axis,
but is introduced by the phase space transformation (\ref{Trafo:x->y}) alone.
To demonstrate this phenomenon visually, we have to take $\epsilon$ not too
small, so we took
$\epsilon = \frac{1}{4}$. Figure \ref{Fig:levelcurves:adiabaticinv} shows a few
level curves
of the adiabatic invariant for $a(\epsilon t) = e^{-\epsilon t}$ and $t$ fixed
at infinity.
The asymmetric effect is clearly present.
\begin{figure}
\begin{center}
\includegraphics[height=8cm]{fig4.eps}
\end{center}
\caption{A few level curves of the adiabatic invariant for $t$ fixed at
infinity}
\label{Fig:levelcurves:adiabaticinv}
\end{figure}
As explained before, we expected to see effects of the slowly decaying
asymmetry in the
neighbourhood of the boundary layer separating the stable and unstable region,
but now it turns out that there are effects ($O(\epsilon)$) close to the origin
too.
\subsection{Approaching the boundary layer}
\label{subsection:Approaching the boundary layer}
We study the approach to the boundary layer, which is an $o(1)$ domain near
the homoclinic orbit and the stable manifold.
More precisely, the boundary layer can be divided into three regions
(see figure \ref{Fig:structure boundary layer}). The first region consists of
the
phase points which are between $O(\exp(-\frac{1}{\epsilon}))$ and $o(1)$ inside
the homoclinic orbit of the unperturbed system.
It is in this region that the averaging technique slowly loses its validity, as
explained in
section \ref{Argument: Averaging breaks down}. We will call this region the
\emph{$o(1)$ boundary layer}.
The second region consists of the phase points which are within an
$O(\exp(-\frac{1}{\epsilon}))$ neighbourhood of
the boundary of the stable region. Orbits starting in these points will pass
the saddle point $(y_1, y_2) = (1, 0)$ on at least a $\frac{1}{\epsilon}$
timescale (which requires special attention), after which they will enter the
third region.
We will call this region the \emph{$O(\exp(-\frac{1}{\epsilon}))$ boundary
layer}.
The third region consists of the remaining phase points in the boundary layer,
which is a strip with an $O(\epsilon)$ width. Orbits starting in this region
will enter the first region
on an $O(1)$ timescale, which allows us to use the unperturbed orbits inside
this region.
We will call this region the \emph{$O(\epsilon)$ boundary layer}.
The \emph{inner region}, finally, consists of the phase points inside the
stable region but \emph{outside} the boundary layer.
\begin{figure}
\begin{center}
\includegraphics[height=8cm]{boundary_layer.eps}
\end{center}
\caption{The structure of the boundary layer}
\label{Fig:structure boundary layer}
\end{figure}
Using the same approach as in the previous subsection, we are able to study the
adiabatic
invariant everywhere in the inner region and the $o(1)$ boundary layer. The
general idea is to expand the averaged equation
around a certain value of $\bar{E}$, in the neighbourhood of which we want the
study the
adiabatic invariant. This can be done to any desired precision. For low order
expansions,
it is possible to integrate the resulting equation explicitly. For high orders,
one has to use
numerical methods.
Approaching the boundary layer, there are two more special values of $\bar{E}$
which we will
study now, knowing the value of $\bar{E}$ corresponding to the maximum of
figure
(\ref{Fig:dependence:E}) and the maximum value $\bar{E} = \frac{1}{6}$.
The first special value can be computed numerically, giving ${\bar{E}}_{max} =
0.1526396\ldots$.
Expanding the averaged equation again to second order around ${\bar{E}}_{max}$,
we arrive at
\begin{equation}
\frac{\mathrm{d}\bar{E}}{\mathrm{d}\tau} = -c_1 \epsilon + c_2 \epsilon
{(\bar{E} - {\bar{E}}_{max})}^2
\end{equation}
with $c_1 = 0.2483204\ldots$ and $c_2 = 64.73966\ldots$, which is easy to solve
giving
\begin{equation}
\bar{E} - {\bar{E}}_{max} = \sqrt{\frac{c_1}{c_2}} \tanh \left( \sqrt{c_1
c_2} (-\epsilon \tau + I_{{\bar{E}}_{max}}) \right)
\end{equation}
where $I_{{\bar{E}}_{max}}$ is an integration constant determined by the
initial condition.
Solving $I_{{\bar{E}}_{max}}$ from this equation, we arrive again at the
adiabatic invariant:
\begin{equation}
I_{{\bar{E}}_{max}} = \frac{1}{\sqrt{c_1 c_2}} \mathrm{artanh} \left(
\sqrt{\frac{c_2}{c_1}} (\bar{E} - {\bar{E}}_{max}) \right) + \epsilon \tau
\end{equation}
Note that these equations hold only in an $O({\epsilon}^{1/3})$ neighbourhood
of ${\bar{E}}_{max}$, and therefore on an ${\epsilon}^{-2/3}$ timescale. For
instance, it does not make sense to take the limit $\tau \rightarrow \infty$.
Although this limit does exist, its value is obviously wrong.
Therefore the $\tanh$ should be regarded only as the first non-trivial
correction to the
linear time evolution of the adiabatic invariant around ${\bar{E}}_{max}$.
The second special value of $\bar{E}$, $\frac{1}{6}$, is much more interesting
and much more tricky, since it lies outside the domain of validity of the
averaging process. We can however still expand the averaged equation around
this value, because the part of the boundary layer inside the homoclinic orbit
of the
unperturbed system ($O(\exp(-\frac{1}{\epsilon}))$) is
much smaller than the domain of validity of the expansion
$O({\epsilon}^{1/2})$.
Therefore, we are allowed to use the results of this expansion, but only
\emph{outside} the boundary layer. Indeed we will see that the results of
this expansion inside the boundary layer are not correct.
Expanding the averaged equation around $\frac{1}{6}$ is not simple, because
the hypergeometric function has an infinite derivative at this point, and the
elliptic integral (the period) is unbounded at this point.
We break up this calculation by expanding equation (\ref{integral:y2:over:y1})
and equation (\ref{period:unperturbed:orbit}) separately.
After a straightforward calculation, we arrive at the following expansions:
\begin{equation}
\label{expansions:E:1/6:int}
\begin{aligned}
\int{y_2}dy_1 = &\frac{6}{5} - 72 \left( 1 - \log \left( \frac{\frac{1}{6}
- \bar{E}}{72} \right) \right) \left( \frac{\frac{1}{6} - \bar{E}}{72}
\right) \\
&+ O\left( \log \left( \frac{1}{6} - \bar{E} \right) {( \frac{1}{6} -
\bar{E} )}^2 \right)
\end{aligned}
\end{equation}
\begin{equation}
\label{expansions:E:1/6:per}
\begin{aligned}
\mathrm{period} = &- \log \left( \frac{\frac{1}{6} - \bar{E}}{72} \right)
-12 \left( 26 + 5 \log \left( \frac{\frac{1}{6} - \bar{E}}{72} \right)
\right) \left( \frac{\frac{1}{6} - \bar{E}}{72} \right) \\
&+ O\left( \log \left( \frac{1}{6} - \bar{E} \right) {( \frac{1}{6} -
\bar{E} )}^2 \right)
\end{aligned}
\end{equation}
Substituting these two expansions, we obtain for the averaged equation (with
$O\left( \epsilon {( \frac{1}{6} - \bar{E} )}^2 \right)$ terms neglected)
\begin{equation}
\label{avg:equation:E:1/6}
\frac{\mathrm{d}\bar{E}}{\mathrm{d}\tau} = \epsilon \frac{12}{5 \log \left(
\frac{\frac{1}{6} - \bar{E}}{72} \right)} + 144 \epsilon \left( 1 -
\frac{2}{\log \left( \frac{\frac{1}{6} - \bar{E}}{72} \right)} - \frac{26}{5
\log^2 \left( \frac{\frac{1}{6} - \bar{E}}{72} \right)} \right) \left(
\frac{\frac{1}{6} - \bar{E}}{72} \right)
\end{equation}
This equation is too complicated to be solved analytically. However, if we
neglect the $O\left( \epsilon ( \frac{1}{6} - \bar{E} ) \right)$ term too, it
is again possible to calculate the adiabatic invariant explicitly:
\begin{equation}
\label{AI:E:1/6}
I_{{\bar{E}}_{1/6}} = \left( \frac{\frac{1}{6} - \bar{E}}{72} \right) \log
\left( \frac{\frac{1}{6} - \bar{E}}{72} \right) - \left( \frac{\frac{1}{6} -
\bar{E}}{72} \right) + \frac{1}{30} \epsilon \tau
\end{equation}
Note that this adiabatic invariant is only valid on an
$\frac{1}{\sqrt{\epsilon}}$ timescale, since $\frac{1}{6} - \bar{E}$ will
become $O(\sqrt{\epsilon})$ on this timescale, causing an extra error of
$O(\epsilon)$.
At this point we are able to make some important remarks:
\begin{itemize}
\item{\emph{Every} orbit starting inside the $o(1)$ boundary layer will
collapse onto the attracting focus $(y_1, y_2) = (0, 0)$ on an
$\frac{1}{\epsilon}$ timescale, \emph{independent} of the initial distance from
the homoclinic orbit (collapsing onto the origin in the $(y_1, y_2)$ plane is
equivalent to circling around the origin in the $(x_1, x_2)$ plane). This
follows
directly from the adiabatic invariant (\ref{AI:E:1/6}), which forces the orbits
away from the boundary layer on an $\frac{1}{\epsilon}$ timescale.}
\item{The averaging process breaks down in the small strip between the $o(1)$
boundary layer and the homoclinic orbit of the unperturbed system, like we
expected it to. If the averaging process would be valid there too, \emph{every}
orbit starting there would collapse onto $(y_1, y_2) = (0, 0)$ on an
$\frac{1}{\epsilon}$ timescale, which would imply that all these orbits stay
within a certain bounded neighbourhood of the origin in the $(x_1, x_2)$
plane.
This cannot be true of course, because an orbit starting very close to the
saddle point $(x_1, x_2) = (1, 0)$ inside the homoclinic orbit, will end up
arbitrary far away from the origin in the $(x_1, x_2)$ plane.}
\item{The leading order behaviour of the period near the homoclinic orbit is
given by $- \log \left( \frac{\frac{1}{6} - \bar{E}}{72} \right)$. This is the
cause of the break-down of the averaging process, since averaging is only valid
if the period is $o(\frac{1}{\epsilon})$.}
\end{itemize}
\section{The boundary layer}
\label{Section: The boundary layer}
After the previous study of the major part of the stable region, we will turn
our attention to
the remaining part of the boundary layer. Since the $o(1)$ boundary layer is
covered by the previous
section, we only have to study the $O(\epsilon)$ and
$O(\exp(-\frac{1}{\epsilon}))$ boundary layers.
As we explained in the previous section, we cannot use the theory of averaging
for this study.
We treat the $O(\epsilon)$ and $O(\exp(-\frac{1}{\epsilon}))$ boundary layers
simultaneously. The
only difference between them is that orbits starting inside the
$O(\exp(-\frac{1}{\epsilon}))$
boundary layer will pass the the saddle point $(y_1, y_2) = (1, 0)$ on at least
a $\frac{1}{\epsilon}$
timescale, which results in an arbitrary large circle in the $(x_1, x_2)$ plane
as t tends to infinity. However this does
not require a separate treatment.
It is important to note that orbits starting inside the $O(\epsilon)$ boundary
layer will remain
within an $O(1)$ neighbourhood of the origin in the $(x_1, x_2)$ plane. So,
although the
$O(\epsilon)$ boundary layer appears to be larger than the
$O(\exp(-\frac{1}{\epsilon}))$ boundary
layer, it is in fact much smaller, because the latter has to fill up the rest
of the $(x_1, x_2)$ phase space.
To study the boundary layer, we can use the same method we used to compute the
position of the
boundary of the stable region, because the orbits in the boundary layer are
close to the
homoclinic orbit of the unperturbed problem. So, to calculate the orbits to
$O(\epsilon)$
precision, we are allowed to substitute expressions for the homoclinic orbit
into the $O(\epsilon)$
contributions to the dynamics.
This way we get again a two stage process. The first stage is governed by the
homoclinic orbit
of the unperturbed system. After the orbit has entered the domain of validity
of the averaging
process, the orbit collapses onto the origin on a $\frac{1}{\epsilon}$
timescale.
Since the existence of an adiabatic invariant was of great help in our study of
the inner region,
we prefer to extend that approach to the boundary layer. We expect an adiabatic
invariant to be
present in the boundary layer too, because we are studying a Hamiltonian system
which depends
adiabatically on time. Finding this adiabatic invariant is generally very hard
in regions where the
unperturbed system has non-periodic solutions (in our case, outside the
homoclinic orbit).
The straightforward way to find the adiabatic invariant is to perturb the
energy of the unperturbed
system in such a way that its time-derivative becomes $O(\epsilon^2)$. So we
are looking for an adiabatic invariant of the form
\begin{equation}
\label{AI:bl}
I_{bl}(y_1, y_2, \epsilon t) = E(\epsilon = 0) + \epsilon g(E(\epsilon = 0),
y_1, \epsilon t)
\end{equation}
where $E(\epsilon = 0)$ is given by (\ref{Def:integral:eps=0}).
By demanding that the time-derivative of $I_{bl}$ has a zero $O(\epsilon)$
contribution, one
normally arrives at a first order linear PDE for the function g. With a little
bit of foresight, we
choose the first argument to be orthogonal to the characteristic lines of the
PDE, which is why we
arrive at a first order linear ODE for the function g.
By using Gradshteyn and Ryzhik~\cite{G&R} intensively we derived this explicit
expression for the function g:
\begin{equation}
\begin{aligned}
g(E(\epsilon &= 0), y_1, \epsilon t) = - \frac{8}{15} \sqrt{\frac{2}{3}}
\frac{a'(\epsilon t)}{a(\epsilon t)} \times \\
&\times \left\{ \left( - b \xi + \frac{3}{2} \xi^3 \right)
\sqrt{{(\xi^2-b)}^2+c^2} +
(b^3 + b c^2) I_1 -\frac{9}{4} I_2 \right\} \\
I_1 &= \frac{1}{2} {(b^2 + c^2)}^{-1/4} \sideset{_2}{_1}F(\alpha, r) \\
I_2 &= \frac{1}{2} {(b^2 + c^2)}^{1/4} \left( \sideset{_2}{_1}F(\alpha, r)
- 2 E(\alpha, r) \right)
+ \frac{\sqrt{{(\xi^2-b)}^2+c^2}}{\xi + \sqrt{b^2 + c^2} \xi^{-1}} \\
\alpha &= \arccos \left\{ \frac{ \sqrt{b^2 + c^2} - \xi^2 }{ \sqrt{b^2 +
c^2} + \xi^2 } \right\} \\
r &= \frac{1}{2} \left( 1 + \frac{b}{\sqrt{b^2 + c^2}} \right)
\end{aligned}
\end{equation}
Note that $E(\alpha, r)$ is the elliptic integral of the second kind and not
the energy. The real numbers $a$, $b$ and $c$ are related to the roots of a
third order polynomial in this way
\begin{equation}
2 E(\epsilon = 0) - y_1^2 + \frac{2}{3} y_1^3 = \frac{2}{3} (y_1 - a)({(y_1 -
a - b)}^2 + c^2) \qquad \forall y_1
\end{equation}
So $a$, $b$ and $c$ are functions of $E(\epsilon = 0)$, with $a < 0$, $b \geq
0$ and $c \geq 0$.
We want to make four remarks with respect to this formula:
\begin{itemize}
\item We have derived the adiabatic invariant outside the homoclinic orbit
(but inside the stable
region) of the unperturbed
system. It is however \emph{not} possible to calculate the adiabatic
invariant in the
inner region using the same procedure, since the characteristic lines
are closed curves
in the inner region, which prohibits the PDE to have a solution.
Indeed, the adiabatic
invariant we found previously for the inner region has an $O(\epsilon)$
time-derivative.
\item $I_{bl}$ determines the dynamics inside the boundary layer completely.
This follows
easily from $d(I_{bl}) = 0$.
\item $I_{bl}$ is symmetric in $y_2$. Transforming back to the $(x_1, x_2)$
plane introduces
again the cross-terms $x_1 x_2$ in the adiabatic invariant which do not
vanish for $t$
going to infinity.
\item We now have an adiabatic invariant throughout the entire stable region,
with the exception
of the very thin ($O(\exp(-\frac{1}{\epsilon}))$) region between the
homoclinic orbit of the
unperturbed system and the $o(1)$ boundary layer. This is not a
problem, since we can
approximate the dynamics inside this strip with transversal orbits
which introduces only an
$O(\frac{1}{\epsilon} \exp(-\frac{1}{\epsilon}))$ error. This trick
solves the problem of matching the two adiabatic invariants at the same
time.
\end{itemize}
We would like to visualize the dynamics going on inside the boundary layer.
Density
functions are not very useful for this, since our system is Hamiltonian which
implies area
conservation. This is well known for autonomous and time-periodic Hamiltonian
systems. To
prove it for general time-dependent systems, one introduces a new independent
variable equal
to the time (making the system autonomous). Applying Liouville's theorem proves
the desired
result.
Note that the conservation of area implies that the area of the ``tongue'' of
the boundary
layer is infinite, since it has to fill up the entire $(x_1, x_2)$ phase space
in the end.
So our ``thin'' boundary layer is in fact the largest part of the stable
region.
To study the dynamics inside the boundary layer, we therefore choose to look at
the evolution
of the rectangular box around the
homoclinic orbit of the unperturbed system, as depicted in figure
\ref{Fig:evolution:box}.
\begin{figure}
\begin{center}
\includegraphics[height=8cm]{fig5.eps}
\end{center}
\caption{The rectangular box around the homoclinic orbit of the unperturbed
system.}
\label{Fig:evolution:box}
\end{figure}
Since all orbits starting inside the box will remain inside the (evoluted) box,
we only have
to study the boundary of the box. Moreover, we only have to study those points
of the boundary
lying inside the stable region, since all other points clear off to infinity on
an $O(1)$
timescale. Therefore we only have to study the bottom boundary (b) of the box.
So by studying only a very limited set of phase space, we will gain information
about all
orbits starting inside the box, i.e. both inside the domain of validity of
averaging and
inside the boundary layer.
For numerical reasons, we followed the evolution of the bottom boundary of a
different (but similar) box in the $(x_1, x_2)$ phase space, namely the
straight line between $(x_1, x_2) = (0, -2.5)$ and $(x_1, x_2) = (5, -2.5)$.
The numerical results are shown
in figure \ref{Fig:evolution}. We took $\epsilon = 0.1$ which forced us to take
steps along
the boundary as small as $10^{-14}$ to generate the last sub-figure. This is
due to the fact
that the most interesting dynamics takes place in an
$O(\exp(-\frac{1}{\epsilon}))$
neighbourhood of the boundary of the stable region.
\begin{figure}
\begin{center}
\includegraphics[width=12.5cm]{fig6.eps}
\end{center}
\caption{The evolution of a straight line (part of the bottom boundary b)
crossing the boundary of the stable region.}
\label{Fig:evolution}
\end{figure}
The open area around the origin is the domain of validity of averaging. This is
the part of
phase space where the level curves of the adiabatic invariant (figure
\ref{Fig:levelcurves:adiabaticinv}) live.
It is also clear to see the instantaneous saddle point moving from $(1,0)$ to
infinity.
The points connecting the instantaneous saddle point with the domain of
validity of averaging
have started very close ($O(\exp(-\frac{1}{\epsilon}))$) to the boundary of the
stable region,
passed the saddle point during a time-interval of $O(\frac{1}{\epsilon})$,
after which they
entered the domain of validity of averaging (in the $(y_1, y_2)$ phase space).
The effect in the $(x_1, x_2)$ phase space is that the orbits end up circling
around the
origin outside the part where the level curves of the adiabatic invariant
(figure \ref{Fig:levelcurves:adiabaticinv}) live.
The closer an orbit starts to the boundary of the stable region, the larger the
radius of the
circle it describes in the end.
The effect of the area conservation is also nicely visible. Since the starting
box has a finite
area, the area inside the spiral must be finite too, which makes the spiral
very thin.
Note that from $t = 4.7$ on the curve going to infinity actually consists of
two very close curves.\\
Note also that the remaining (major) part of phase space has to be filled by
the tail of the
``small'' tongue of the boundary layer which lies outside the box.
\section{Concluding remarks}
It is surprising that it is possible to give a fairly complete treatment of
system (\ref{Sys:x}) which describes the evolution of a simple system with an
asymmetric potential to a symmetric potential. The most remarkable result is
that in the evolution towards symmetry as time tends to infinity, traces of the
asymmetric past can be recognized in the solutions.
System (\ref{Sys:x}) is just a metaphor for simplified models with two degrees
of freedom which exhibit evolution from asymmetry towards symmetry. In a
forthcoming paper we shall discuss such higher dimensional problems using
basically the same methods.
In the discussion of the validity of the averaged equation, we have assumed
that the reader is familiar with the proof of the standard averaging theorems
(see for instance Sanders and Verhulst~\cite{S&V}).
In particular, we have used the straightforward extension of those theorems to
periods depending on $\epsilon$: by rescaling the time-variable it is easily
shown that averaging produces $O(\epsilon T(\epsilon))$ approximations, valid
on a $\frac{1}{\epsilon}$ timescale, as long as $\epsilon T(\epsilon)$ is
$o(1)$.
\label{Argument: Averaging breaks down}
|
\section{Appendix}
\begin{eqnarray}
d_{N\pi}(y)={g^2_{NN\pi}\over 16\pi^2}\int_0^\infty dk^2_\perp
{|G_{N\pi}(y,k^2_\perp|^2\over y^2(1-y)}
{m^2_N(1-y)^2-k^2_\perp\over [m_N^2-M^2_{N\pi}(y,k^2_\perp)]^2},
\end{eqnarray}
\begin{eqnarray}
d_{\Delta\pi}(y)&=&{g^2_{N\Delta\pi}\over 96\pi^2}
\int_0^\infty dk^2_\perp
{|G_{\Delta\pi}(y,k^2_\perp)|^2\over y^4(1-y)m_\Delta^2} \nonumber \\*
&\times& {[(ym_N+m_\Delta)^2+k^2_\perp]
[(y^2m_N^2-m_\Delta^2)^2+8ym_Nm_\Delta k^2_\perp-k^4_\perp]
\over [m_N^2-M^2_{\Delta\pi}(y,k^2_\perp)]^2}.
\end{eqnarray}
and
\begin{eqnarray}
d_{int}(y)&=&{g_{N\Delta\pi} g_{NN\pi}\over 16\sqrt 6\pi^2}
\int_0^\infty dk^2_\perp
{G_{\Delta\pi}(y,k^2_\perp)G_{N\pi}(y,k^2_\perp) \over y^3(1-y)m_\Delta}
\nonumber \\*
&\times&\left\{\frac{-m_N(1-y)(ym_N+m_\Delta)^2(ym_N-m_\Delta)}
{[m_N^2-M^2_{\Delta\pi}(y,k^2_\perp)]
[m_N^2-M^2_{N\pi}(y,k^2_\perp)]}\right. \nonumber \\*
& & \left. + \frac{(2m_\Delta^2+(3y-2)m_Nm_\Delta-ym_N^2)k^2_\perp-k_\perp^4}
{[m_N^2-M^2_{\Delta\pi}(y,k^2_\perp)]
[m_N^2-M^2_{N\pi}(y,k^2_\perp)]}\right\}.
\end{eqnarray}
\addcontentsline{toc}{chapter}{\protect\numberline{}{References}}
|
\section{Introduction}
Since Cheng ~\cite{cheng} showed that there is a
factor-of-two discrepancy between
the empirical data for the pion-nucleon sigma term ($\Sigma_{\pi N}$) and
the naive estimates of the $\sigma$-term from the mass spectrum,
there have been a great deal of discussions and
disputes about the $\Sigma_{\pi N}$ and $\sigma$ term
(see Ref.~\cite{jk,gls} and references therein).
Donoghue and Nappi~\cite{dn} suggested that the discrepancy
is due to the presence of strange quarks in the nucleon, {\em i.e.}
$\langle N| \bar{s}s |N\rangle\neq 0$ and showed that
$\langle N| \bar{s}s |N\rangle$ contributes almost $30\%$ to the
quark condensate in the nucleon, making use of the Skyrme model
and bag model. At the first thought, it seems to be reasonable,
since Cheng used the Zweig rule, {\em i.e.} neglected
$\langle N| \bar{s}s |N\rangle$.
However, one serious question arises: a large fraction of the
nucleon mass then stems from strange quarks if one follows
Ref.~\cite{dn}, which contradicts the quark model.
Another assumption was that the ratio $m_s/\bar{m}$ is off
by a factor of two, which means that the first order
perturbation theory collapses. However, this kind of
suggestion would lead to a breakdown of the
Gell-Mann-Okubo mass formula which predict the masses of
hadrons in a few percent.
Motivated by these contradictions,
Gasser, Leutwyler and Sainio~\cite{gls} recently
reanalysed the $\sigma$ term prudently, taking advantage
of newly accumulated and better $\pi N$ scattering data
and considering the strong $t$-dependence of the scalar
form factor $\sigma(t)$
($\sigma(2m^{2}_{\pi})-\sigma (0)\simeq 15 \;\mbox{MeV}$).
The results of Ref.~\cite{gls} were $\sigma = 45 \pm 8 \mbox{MeV}$
and $\Sigma \simeq 60 \mbox{MeV}$.
The $y=2\langle N| \bar{s}s |N\rangle/\langle N|
\bar{u}u + \bar{d}d |N\rangle$, a share
of $\langle N| \bar{s}s |N\rangle$ in the $\sigma$ term,
was about 0.2, so that the corresponding contribution of the
term $\langle N| \bar{s}s |N\rangle$ to the nucleon mass was
about $130\;\mbox{MeV}$.
In the meanwhile, the efforts to understand the $\sigma$ term puzzle
theoretically have continued~\cite{bft,bass,sw}. However, the bone of
contention still lies in the role of strange quarks, more specifically
the contribution of the $\langle N| \bar{s}s |N\rangle$ to the
$\sigma$ term. Recently, several works insist that
there is no need to introduce a portion of strange quarks to
explain the $\sigma$ term discrepancy.
Bass~\cite{bass} proposed that based on the
Gribov confinement the value of the $\sigma$
term can be explained without need to invoke large strangeness
content of the nucleon. Ball, Forte and Tigg~\cite{bft} also
suggested that with the correct understanding of the baryon matrix
element the $\sigma$ term (identified with
$\sigma_8 =\bar{m} \langle N| \bar{u}u +\bar{d}d -2\bar{s}s
| N \rangle$ ) can be reproduced without
violating the Zweig rule. Hence, following these arguments,
strange quarks do not contribute to the nucleon mass.
Though it should be small, it is still important to consider
the contribution of strange quarks to the $\sigma$ term,
in line with recent experiments indicating the fact
that strange quarks might play an important role of explaining
the properties of the nucleon~\cite{emc,bnl}.
It is the object of the present work to study the strangeness
contribution to the $\sigma$ term in the framework of the
semi-bosonized SU(3) Nambu-Jona-Lasinio soliton model (often called
as the chiral quark soliton model).
In our model, the nucleon is understood explicitly as
$N_c$ valence quarks coupled to the polarized Dirac sea
bound by a non-trivial chiral mean field configuration.
The proper quantum numbers of the nucleon can be acquired by
the semiclassical quantization~\cite{dpp,anw} performed
via integrating over the zero-mode fluctuations of the pion field
around the saddle point. It allows the nucleon to carry proper
quantum numbers such as spins and isospins.
The SU(3) NJL soliton model has a merit in that it interpolates between
the nonrelativistic naive quark model and the Skyrme model.
It enables us to study the interplay between these two
different models~\cite{PraBlGo}.
The model is quite successful in describing the
static properties of the baryons and their form factors
\cite{betal,bpg,kbpg}.
The outline of the paper is as follows:
In the next section, we sketch the basic formalism for
the scalar form factor in SU(3) NJL soliton the model.
In section 3, we present the numerical results and
discuss about them.
In section 4, we summarize the present work and remark the conclusion.
\section{Formalism}
The scalar form factor $\sigma (t)$ is defined as a condensate of
$u$ and $d$ quarks in the nucleon:
\begin{equation}
\sigma (t) \;=\; \bar{m}
\langle N(p') |\bar{u}u + \bar{d}{d} | N(p) \rangle
\label{Eq:sigma1}
\end{equation}
with $\bar{m}=(m_u+m_d)/2\simeq6\;\mbox{MeV}$.
The $t$ denotes the square of the momentum transfer.
Our model is characterized by a
low--momenta QCD partition function in Euclidean space
given by the functional integral over pseudoscalar meson
and quark fields:
\begin{equation}
{\cal Z} \; = \; \int {\cal D} \Psi {\cal D} \Psi^{\dagger}
{\cal D} \pi^A \exp{\left(-\int d^4x\Psi^\dagger iD\Psi\right)}
\label{Eq:func}
\end{equation}
where
\begin{equation}
iD \;=\; \beta(-i\rlap{/}{\partial}\;+\;
MU^{\gamma_5}\;+\;\hat{m})
,\;\;U^{\gamma_5}=e^{i\pi^a\lambda^a\gamma_5}.
\end{equation}
$\lambda^a$ are SU(3) Gell-Mann matrices normalized as
$\mbox{Tr}{\lambda^a\lambda^b} = 2\delta^{ab}$. The $\hat{m}$
denotes the current quark mass matrix for which
we take the form $diag(m_u,m_d,m_s)$, where
$m_u, m_d$ and $m_s$ are the corresponding current quark masses
of the {\em up}, {\em down} and {\em strange} quark, respectively.
Here, we assume that isospin symmetry is not broken, {\em i.e.}
$m_u=m_d=\bar{m}$.
The $M$ stands for the momentum--dependent dynamical mass arising
from the spontaneous chiral symmetry breaking.
The momentum--dependence of the $M$ introduces the ultra--violet
cut--off. However, we shall regard it as a constant for simplicity.
Instead, we employ a simple proper--time regularization.
The differential operator $iD$ is expressed in Euclidean space in terms of
the Euclidean time derivative $\partial_\tau$, the Dirac one-particle
Hamiltonian $H(U)$ and symmetry breaking part~\cite{sbg}:
\begin{equation}
iD\;=\;\partial_\tau \;+\; H(U) \;+\; h_{sb}
\end{equation}
with
\begin{equation}
H(U)\;=\;\frac{\vec{\alpha}\cdot \nabla}{i} \;+\; \beta M_u U
\;+\; \beta \bar{m}{\bf 1},\;\;\; h_{sb} \;=\;\beta \mu_0 {\bf 1}
\;+\;\beta \mu_8 \lambda_8.
\end{equation}
Here, we have made the famous embedding Ansatz for the pseudoscalar
fields $U^{\gamma^5}$ and
$U$ is expressed by
\begin{equation}
U\;=\; \left(\begin{array}{cc} U_{0} & 0 \\
0 & 1
\end{array} \right).
\end{equation}
The $U_{0}$ expresses the SU(2) chiral background field
$U_0=\exp{i[\vec{n}\cdot\vec{\tau} P(r)]}$ with the hedgehog Ansatz.
$P(r)$ denotes the profile function with proper boundary conditions.
$\mu_0$ and $\mu_8$ are defined by
$\mu_0 = (M_s - M_u) / 3$ and $\mu_8 = -(M_s - M_u) / \sqrt{3}$.
$M_s$ and $M_u$ are constituent quark masses of the $s$ and $u$ quarks
respectively. The $M_u$ is used as an input parameter, while
the $M_s$ is determined by the gap equation~\cite{sbg}.
The current strange quark mass $m_s$ is also settled in the same way.
We treat the explicit symmetry breaking term $h_{sb}$ perturbatively.
The hadronic matrix elements of the $\pi\mbox{N}\; \sigma$--term
is related to the correlation function
\begin{equation}
\sigma(t)\;\smash{\mathop{\sim}\limits_{T\rightarrow \infty}}\;
\langle 0 | J_N (\vec{x}, \frac{T}{2})\hat{\sigma}
J^{\dagger}_{N} (\vec{y}, -\frac{T}{2}) | 0 \rangle
\label{Eq:sigma}
\end{equation}
at large Euclidean time $T$. $\hat{\sigma}$ is the quark operator
for the $\sigma$ term, defined by
$\hat{\sigma}=\bar{m}(\bar{u}u+\bar{d}d)$.
$J_N$ is the nucleon current
constructed from $N_c$ quark fields~\cite{dpp}
\begin{equation}
J_N(x)\;=\; \frac{1}{N_c !} \epsilon_{i_1 \cdots i_{N_c}}
\Gamma^{\alpha_1 \cdots
\alpha_{N_c}}_{JJ_3TT_3Y}\psi_{\alpha_1i_1}(x)
\cdots \psi_{\alpha_{N_c}i_{N_c}}(x).
\end{equation}
$\alpha_1 \cdots\alpha_{N_c}$ denote spin--flavor indices, while
$i_1 \cdots i_{N_c}$ designate color indices. The matrices
$\Gamma^{\alpha_1 \cdots\alpha_{N_c}}_{JJ_3TT_3Y}$ are taken to endow
the corresponding current with the quantum numbers $JJ_3TT_3Y$.
The $J^{\dagger}_{B}$ plays the role of creating the baryon state.
The integral over the quark fields are trivial.
The integral over the pseudo-Goldstone boson fields can be performed
by the saddle point method in the large $N_c$ limit.
In order to find the quantum $1/N_c$ corrections,
it is important to take into account the small oscillations of the
pseudo-Goldstone bosons around the saddle point and the zero modes.
The zero modes are taken into account by the
soliton expressed by $\tilde{U} (\vec{x},x_4)
=A(x_4)U(\vec{x} - \vec{Z})A^{\dagger} (x_4)$
with an SU(3) unitary matrix $A(t)$.
Hence, the collective action $S_{eff}$ becomes
\begin{eqnarray}
\tilde{S}_{eff} & = & -N_c {\rm Sp} \log{(iD)} \nonumber \\
& = & -N_c \mbox{Sp}
\log{\left [ \partial_\tau \;+\; H(\tilde{U})
\;+\; A^{\dagger} (x_4) \dot{A}(x_4)
\;-\; i \beta \dot{\vec{Z}} \cdot \nabla \right. }
\nonumber \\
& & \; +\; \left. A^{\dagger}(x_4) h_{sb} A(x_4)
\;-\; \xi (y) \beta A^{\dagger}(x_4) \frac{1}{\sqrt{3}}
(\sqrt{2}\lambda_0 + \lambda_8) A(x_4) \right]
\end{eqnarray}
with the angular velocity $A^\dagger (x_4) \dot{A} (x_4) = i\Omega_E
=i\Omega^{a}_{E} \lambda^a /2$. $\mbox{Sp}$ denotes the
functional trace.
The $\xi$ stands for
the external scalar field, with regard to which
we make a functional derivative so as to obtain the sigma form factor:
\begin{eqnarray}
\sigma(t) &=& -N_c
\frac{\delta }{\delta \xi(z)} \mbox{Sp}
\log{\left \{ \partial_\tau \;+\; H(\tilde{U})
\;+\; A^{\dagger} (x_4) \dot{A}(x_4)
\;-\; i \beta \dot{\vec{Z}} \cdot \nabla \right. }
\nonumber \\
& & \; +\; \left. A^{\dagger}(x_4) h_{sb} A(x_4)
\;-\; \xi (y) \beta A^{\dagger}(x_4) \frac{1}{\sqrt{3}}
(\sqrt{2}\lambda_0 + \lambda_8) A(x_4) \right \}
\end{eqnarray}
It is known that there is the dependence of
the $\sigma$ term on the regularization scheme~\cite{admg}.
However, we want to stress the fact that we have employed
the proper-time regularization and have evaluated
possible physical observables such as mass splittings,
magnetic moments, axial constants and electromagnetic form factors
within the same scheme and same values of input parameters
\footnote{In fact,
we have only one free parameter, {\em i.e.} the constituent
up-quark (down-quark) mass. However, it is more or less fixed
to around $420\;\mbox{MeV}$ by
the mass splitting~\cite{betal}. }.
Hence, we stick to
the proper-time regularization for the $\sigma$ term
and make use of the same input parameters without adjusting.
However, we shall not be here bothered by going through all the
tedious technical details arising from the regularization
(see Ref.~\cite{kpbg} for details).
Having taken into account the rotational $1/N_c$ corrections and
linear $m_s$ corrections, we arrive at
\begin{eqnarray}
\sigma (t) & = & \Sigma_{SU(2)} (t) \langle 2 \;+\; D^{(8)}_{88}(A)
\rangle_N \nonumber \\
& + & \frac{2\bar{m}}{\sqrt{3}I_1} {\cal K}_1 (t)
\langle D^{(8)}_{8i}(A)R_i \rangle_N
\; + \; \frac{2\bar{m}}{\sqrt{3} I_2} {\cal K}_2 (t)
\langle D^{(8)}_{8p}(A)R_p \rangle_N \nonumber \\
& - & \frac{4\bar{m}\mu_8}{\sqrt{3}}
\left[{\cal N}_1(t) - {\cal K}_1 (t) \frac{K_1}{I_1}\right]
\langle D^{(8)}_{8i}(A)D^{(8)}_{8i}(A) \rangle_N
\nonumber \\
& - & \frac{4\bar{m}\mu_8}{\sqrt{3}}
\left[{\cal N}_2(t) - {\cal K}_2 (t) \frac{K_2}{I_2}\right]
\langle
D^{(8)}_{8p}(A)D^{(8)}_{8p}(A) \rangle_N \nonumber \\
&-& \frac{4\bar{m}\mu_8}{3\sqrt{3}}
{\cal N}_0 (t)
\langle D^{(8)}_{88}(A) (D^{(8)}_{88}(A) + 1) \rangle_N
\; - \; \frac{8\bar{m}\mu_0}{3} {\cal N}_0 (t),
\label{Eq:sf}
\end{eqnarray}
where
\begin{eqnarray}
\Sigma_{SU(2)} (t) & = & N_c \int d^3x \;j_0 (Qr)
\left [\Psi^{\dagger}_{val} (x) \beta \Psi_{val} (x) \;-\; \sum_n
\frac{1}{2} \mbox{sign}(E_n) {\cal R} (E_n)
\Psi^{\dagger}_{n} (x) \beta \Psi_n (x) \right], \nonumber \\
{\cal K}_1 (t) & = & \frac{N_c}{6} \sum_{n,m}
\int d^3x j_0 (Qr) \int d^3y \left[
\frac{\Psi^{\dagger}_{n} (x) \vec{\tau} \Psi_{val} (x) \cdot
\Psi^{\dagger}_{val} (y) \beta \vec{\tau} \Psi_{n} (y)}
{E_n - E_{val}} \right .
\nonumber \\ & & \hspace{3cm} \;+\; \left . \frac{1}{2}
\Psi^{\dagger}_{n} (x) \vec{\tau} \Psi_{m} (x) \cdot
\Psi^{\dagger}_{m} (y) \beta \vec{\tau} \Psi_{n} (y)
{\cal R}_{\cal M} (E_n, E_m)
\right ],
\nonumber \\
{\cal K}_2 (t) & = & \frac{N_c}{6} \sum_{n, m^{0}}
\int d^3 x \;j_0 (Qr) \int d^3 y
\left [\frac{\Psi^{\dagger}_{m^{0}} (x) \Psi_{val} (x)
\Psi^{\dagger}_{val} (y) \beta \Psi_{m{^0}} (y)}
{E_{m^{0}} - E_{val}} \right .
\nonumber \\ & & \hspace{3cm} \;+\;\left . \frac{1}{2}
\Psi^{\dagger}_{n} (x) \Psi_{m^{0}} (x)
\Psi^{\dagger}_{m^{0}} (y) \beta \Psi_{n} (y)
{\cal R}_{\cal M} (E_n, E_m^{0}) \right ], \nonumber \\
{\cal N}_1 (t) & = & \frac{N_c}{6} \sum_{n,m}
\int \; d^3x j_0 (Qr) \int d^3y \left[
\frac{\Psi^{\dagger}_{n} (x) \beta\vec{\tau} \Psi_{val} (x) \cdot
\Psi^{\dagger}_{val} (y) \beta \vec{\tau} \Psi_{n} (y)}
{E_n - E_{val}} \right .
\nonumber \\ & & \hspace{3cm} \;+\; \left . \frac{1}{2}
\Psi^{\dagger}_{n} (x) \beta \vec{\tau} \Psi_{m} (x) \cdot
\Psi^{\dagger}_{m} (y) \beta \vec{\tau} \Psi_{n} (y)
{\cal R}_{\beta} (E_n, E_m)
\right ],
\nonumber \\
{\cal N}_2 (t) & = & \frac{N_c}{6} \sum_{n, m^{0}}
\int d^3 x \;j_0 (Qr) \int d^3 y
\left [\frac{\Psi^{\dagger}_{m^{0}} (x)\beta \Psi_{val} (x)
\Psi^{\dagger}_{val} (y) \beta \Psi_{m{^0}} (y)}
{E_{m^{0}} - E_{val}} \right .
\nonumber \\ & & \hspace{3cm} \;+\;\left . \frac{1}{2}
\Psi^{\dagger}_{n} (x) \beta \Psi_{m^{0}} (x)
\Psi^{\dagger}_{m^{0}} (y) \beta \Psi_{n} (y)
{\cal R}_{\beta} (E_n, E_m^{0}) \right ], \nonumber \\
{\cal N}_0 (t) &=& \frac{3N_c}{2}
\sum_{n,m}
\int d^3x j_0 (Qr) \int d^3y \left[
\frac{\Psi^{\dagger}_{n} (x) \beta\Psi_{val} (x)
\Psi^{\dagger}_{val} (y) \beta \Psi_{n} (y)}
{E_n - E_{val}} \right .
\nonumber \\ & & \hspace{3cm} \;+\; \left . \frac{1}{2}
\Psi^{\dagger}_{n} (x) \beta \Psi_{m} (x) \cdot
\Psi^{\dagger}_{m} (y) \beta \Psi_{n} (y)
{\cal R}_{\beta} (E_n, E_m)
\right ].
\end{eqnarray}
The subscripts $i$ and $p$ in the collective part are
$i=1,2,3$ and $p=4,5,6,7$, respectively.
$I_i$ and $K_i$ are respectively the moments of inertia and
anomalous moments of inertia~\cite{betal}. When $t\rightarrow 0$,
${\cal K}_i (t) $ become $K_i$. The $\Sigma_{SU(2)}$ corresponds to
the $\pi N$ sigma term in SU(2)~\cite{betal} at $t=0$,
which can be obtained by the Feynman-Hellman theorem
\begin{equation}
\Sigma_{SU(2)} = \left. \bar{m}
\frac{\partial E(\bar{m})}{\partial \bar{m}}\right|_{\bar{m}=0},
\end{equation}
where $E$ stands for the classical soliton energy.
The regularization functions $R (E_n)$, $R_{\cal M} (E_n, E_m)$,
$R_{\beta} (E_n, E_m)$
\footnote{$R_{\cal M} (E_n,E_m)$ is not actually a
regularization function, since ${\cal K}_i$ come from the imaginary
part of the action. It does not depend on the cut-off parameter.}
are defined by
\begin{eqnarray}
{\cal R} (E_n) & = & \int \frac{du}{\sqrt{\pi u}}
\phi (u;\Lambda_i) |E_n| e^{-uE^{2}_{n}},
\nonumber \\
{\cal R}_{\cal M} (E_n, E_m) & = &
\frac{1}{2} \frac{ {\rm sign} (E_n)
- {\rm sign} (E_m)}{E_n - E_m},
\nonumber \\
{\cal R}_{\beta} (E_n, E_m) & = &
\int^{\infty}_{0} \frac{du}{2\sqrt{\pi u}} \phi (u;\Lambda_i)
\frac{E_n e^{-u E^{2}_{n}} - E_m e^{-u E^{2}_{m}}}{E_n - E_m},
\end{eqnarray}
respectively.
The $\langle\rangle_N$ stands for the expectation value of the
Wigner $D$ functions in collective space apanned by $A$.
The expectation values of the $D$ functions can be evaluated by
SU(3) Clebsch-Gordan coefficients found in Refs.~\cite{Swart,Mcnamee}.
With SU(3) symmetry explicitly broken by $m_s$, the collective part is no
longer SU(3)-symmetric. Therefore, the eigenstates of the
hamiltonian are not in a pure octet or decuplet but mixed states.
Since we treat the strange quark mass $m_s$ perturbatively,
we can obtain the mixed SU(3) baryonic states as follows:
\begin{equation}
| 8, N \rangle \;=\; | 8, N \rangle \;+ \;
c^{N}_{\bar{10}} | \bar{10}, N \rangle
\;+\;c^{N}_{27} | 27, N \rangle
\end{equation}
with
\begin{equation}
c^{N}_{\bar{10}} \;=\; \frac{\sqrt{5}}{15}(\bar{\sigma} - r_1)
I_2 m_s,
c^{N}_{27} \;=\; \frac{\sqrt{6}}{75}(3\bar{\sigma} + r_1 - 4r_2)
I_2 m_s.
\label{Eq:g2}
\end{equation}
The constant $\bar{\sigma}$ is related to the $\Sigma_{SU(2)}$
by $\Sigma_{SU(2)} = 2/3 (m_u + m_d)\bar{\sigma}$. $r_i$ denotes
the ratio $K_i / I_i$.
Since the Cheng-Dashen point is out of the physical region,
it is necessary to extrapolate to the region $t>0$.
This can be done by the analytic continuation of the $|\vec{q}|$,
{\em i.e.} $|\vec{q}| \rightarrow i|\vec{q}|$ so that
we may have the positive $t$ up to the Cheng-Dashen point
($t=2m^{2}_{\pi}$). The analytic continuation
above the threshold $t=4m^{2}_{\pi}$ is not valid in our model,
since above this threshold, the correlation between mesonic
clouds is getting important~\cite{phs}.
Hence, in this work, we only evaluate the scalar form factor
from the Cheng-Dashen point to the physical channel
(space-like region: $t<0$).
\section{Numerical Results and Dicussion}
In order to calculate the $\sigma_{\pi N} (t)$
numerically, we take advantage of
the Kahana-Ripka discretized basis~\cite{kr}. Figure 1 shows
the scalar form factor as a function of the constituent quark mass $M
=M_u = M_d$. The $\sigma (t)$ decreases as the $M$ increases,
in particular, below $t=0$. As a result, the difference between
the $\sigma (2m^{2}_{\pi})$ and $\sigma (0)$ changes drastically
when we increase the $M$ from $370\;\mbox{MeV}$ to $450\;\mbox{MeV}$, as
shown in Table 1. We select the $M=420\;\mbox{MeV}$
for the best fit as we did for other observables.
The error bar presented in Fig. 1 stands for the empirical
analysis due to Gasser, Leutwyler,
and Sainio~\cite{gls}, {\em i.e.} $\sigma (0) =45\pm 8 \; \mbox{MeV}$.
Our numerical prediction is in a remarkable agreement with
Ref.~\cite{gls}. It is also interesting to see how the $m_s$ corrections
contribute to the scalar form factor. As shown in Fig. 2, the $m_s$
corrections are very small. At $t=0$, the $m_s$ corrections contribute
to the $\sigma$ term about $2\%$ which is negligible.
However, the $m_s$ corrections
play a significant role of reducing remarkably the large
strangeness contribution $\langle N | \bar{s} s | N \rangle$
arising from the leading term and rotational $1/N_c$
corrections. With the $m_s$ corrections taken into account,
we obtain $y=0.27$ in case of the $M=420\mbox{MeV}$, which
agrees with the empirical value $y\simeq 0.2$~\cite{gls} within
about $30\%$, whereas we
have $y=0.48$ without the $m_s$ corrections. It is already known that
the explicit symmetry breaking term quenchs the
$\langle N | \bar{s} s | N \rangle$~\cite{bjm,hatsuda,kk}.
The difference $\Delta \sigma = \sigma(2m^{2}_{\pi})
-\sigma(0)$ we have obtained is $18.18\mbox{MeV}$. This value is
very close to what Gasser and Leutwyler extracted~\cite{gl},
$\Delta \sigma = 15.2\pm 0.4 \mbox{MeV}$. The tangent of
the scalar form factor at $t=0$ is known to be related to
the scalar square radius. It is almost
two times larger than the electric one, {\em i.e.}
the $\langle r^2\rangle^{S}_{N} \simeq 1.6 \mbox{fm}^2$
while $\langle r^2\rangle^{E}_{N} \simeq 0.74\mbox{fm}^2$.
The prediction of our model for the $\langle r^2\rangle^{S}_{N}$
is $1.5 \mbox{fm}^2$ which is almost the same as
obtained by Gasser and Leutwyler.
It implies that the tail of the
scalar density is of great importance. In Fig. 3 we can find a
long-stretched and strong tail in the sea contribution to the scalar
density. This tail is due to the mesonic clouds arising from the
Dirac sea polarization.
Moreover, the sea contribution in the scalar density is
large, compared with the other densities such as
electromagnetic densities~\cite{wkg,wakamatsu,cggp}.
The other interesting quantities are presented in table 1.
$\sigma_0$ is the condensate of the singlet scalar quark operator
in the nucleon:$\sigma_0 = \bar{m} \langle N |
\bar{u}u+\bar{d}d + \bar{s}s| N \rangle$
$R_s$ is defined by
$R_{s} = \langle N | \bar{s}s | N \rangle /
\langle N | \bar{u}u+\bar{d}d + \bar{s}s| N \rangle$.
\section{Summary and Conclusion}
We have discussed the scalar form factor with related
quantities in the SU(3) NJL soliton model. The results we have obtained
are in a good agreement with empirical data~\cite{gls,gl}.
The reliable strangeness contents of the nucleon in the scalar
channel is obtained by taking into account the $m_s$ corrections,
since they suppress the excess of $\langle N|\bar{s}s | N \rangle$
due to the leading order and rotational $1/N_c$ contributions.
In contrast to Refs.~\cite{bft,bass} suggesting no strangeness
contribution, our model favors $y=0.27$.
The large value of the $\langle r^2\rangle^{S}_{N}$
is caused by the pronounced long ranging tail
which can be identified with the pion and kaon clouds.
\section*{Acknowledgement}
The authors would like to thank M. Polyakov for helpful discussions.
This work has partly been supported by the BMFT, the DFG,
the COSY--Project (J\" ulich) and Department of Energy
grant DE-FG02-88ER40388. One of us (AB) would like to thank
the {\it Alexander von Humboldt Foundation} for a Feodor Lynen grant.
\begin{table}
\caption{The physical quantities related to the scalar form factor.
The empirical data come from ~Ref.[3,17].}
\begin{tabular}{c|c|c|c|c|c|c|c}
$M$ & \multicolumn{2}{c|}{$370$ MeV} & \multicolumn{2}{c|}{$420$ MeV}
& \multicolumn{2}{c|}{$450$ MeV} & Exp \\ \cline{1-7}
$m_s$ [MeV] & 0 & 156.75 & 0 & 148.49 & 0 & 145.35 & \phantom{} \\ \hline
$\sigma_{\pi N}$[MeV] & 43.09 & 44.71 & 40.01 & 40.80
& 38.22 & 38.69 & $45\pm 8$ \\
$\sigma_{0}$[MeV] & 53.25 & 49.25 & 49.58 & 46.24
& 47.37 & 44.35 & \phantom{} \\
$\sigma_{8}$[MeV] & 22.77 & 35.63 & 20.87 & 29.92
& 19.92 & 28.37 & $\phantom{}$ \\
$y$ & 0.47 & 0.20 & 0.48 & 0.27 & 0.48 & 0.29 & $0.2\pm 0.2$ \\
$R_{s}$ & 0.19 & 0.09 & 0.19 & 0.12 & 0.19 & 0.13 & \phantom{} \\
$\Delta\sigma$[MeV] &32.29&33.37&18.36&18.18&14.23&13.84&$15.2\pm0.4$ \\
$ \langle r^2\rangle^{S}_{N}$ &1.94&1.87&1.56&1.50&1.40&1.34&1.6
\end{tabular}
\end{table}
|
\section{Introduction} \label{sec:intro}
Given an initial data set for the gravitational field and any matter
fields present, what can be said of the spacetime evolved from this
initial data?
In the asymptotically flat case, one would like to know such things as
how much gravitational energy is radiated to null infinity, the final
asymptotic state of the system, whether black holes are formed, the
nature of any singularities produced, and whether cosmic censorship is
violated. For example, it is known that the maximal development of
sufficiently weak vacuum initial data is an asymptotically flat
spacetime that is free of singularities and black holes \cite{CK93}.
In this case the gravitational waves are so weak that they cannot
coalesce into a black hole; instead they scatter to infinity. Further
it is known that an initial data set containing a future trapped
surface or a future trapped region must be singular, provided the
null-convergence condition holds \cite{HawkingEllis73,Wald84}. In
these cases, the gravitational field is already sufficiently strong
that collapse is inevitable.
In the cosmological case (spacetimes with compact Cauchy surfaces),
the questions one asks are a bit different as one expects these
spacetimes to be quite singular. In fact, it is known that spacetimes
with compact Cauchy surfaces are singular, provided a genericity
condition and the timelike-convergence condition hold
\cite{HawkingEllis73,Wald84}. So, here one would like to know such
things as the nature of the singularities, if the spacetime has a
finite lifetime (in the sense that there is a global upper bound on
the lengths of all causal curves therein), whether it expands to a
maximal hypersurface and then recollapses or is always expanding
(contracting), and whether cosmic censorship is violated. For
example, it is known that if the initial data surface is contracting
to the future (past), then any development satisfying the
timelike-convergence condition must end within a finite time to the
future (past) \cite{HawkingEllis73,Wald84}. Can more be said about
the behavior of the cosmological spacetimes?
The closed-universe recollapse conjecture asserts that the spacetime
associated with the maximal development of an initial data set with
compact initial data surface expands from an initial singularity to a
maximal hypersurface and then recollapses to a final singularity (all
within a finite time), provided that the spatial topology does not
obstruct the existence of a maximal Cauchy surface (e.g., $S^3$ or
$S^1 \times S^2$) and provided the matter satisfies certain energy and
regularity conditions \cite{MarsdenTipler80,BarrowTipler85,BGT86}. It
has also been conjectured that such spacetimes admit a unique
foliation by constant mean curvature (CMC) Cauchy surfaces with the
mean curvatures taking on all real values. (See, e.g., conjecture~2.3
of \cite{EardleySmarr79} and the weaker conjecture~C2 of
\cite{EardleyMoncrief81}.)\ \ Just what energy conditions the matter
must satisfy is an open problem. However, in the study of the weak
form of this conjecture (which merely asserts that the spacetime has a
finite lifetime), the dominant energy and non-negative pressures
conditions together have proven sufficient for the cases studied
\cite{Burnett95,Burnett91}. More subtle is the problem of what regularity
conditions the matter needs to satisfy. The difficulty here is that
the maximal development of an Einstein-matter initial data set may not
contain a maximal hypersurface because of the development of a
singularity in the matter fields, such as a shell-crossing singularity
in a dust-filled spacetime, before the spacetime has a chance to
develop a maximal hypersurface. While not for certain, it is thought
that those matter fields that do not develop singularities when
evolved in fixed smooth background spacetimes will not lead to the
obstruction of a maximal hypersurface.
Here, we study the maximal development of spherically symmetric
constant mean curvature initial data sets with $S^1 \times S^2$ Cauchy
surfaces and matter consisting of either collisionless particles of
unit mass (whose evolution is described by the Vlasov equation) or a
massless scalar field (whose evolution is described by the massless
wave equation). It has already been established that if the mean
curvature is zero on the initial data surface, i.e., it is a maximal
hypersurface, then its maximal evolution admits a foliation by CMC
Cauchy surfaces with the mean curvature taking on all real values
\cite{Rendall95}. Further, it is known that if the mean curvature is
negative (positive) then the initial data can be evolved at least to
the extent that the spacetime can be foliated by CMC spatial
hypersurfaces taking on all negative (positive) values
\cite{Rendall95}. Left unresolved was whether the maximal evolution
in the latter two cases actually contains a maximal spatial
hypersurface and, hence, can be foliated by CMC hypersurfaces taking
on all real values. The nonexistence of a maximal spatial
hypersurface would be reasonable if such spacetimes could expand
(contract) indefinitely, however, it is known that these spacetimes
have finite lifetimes \cite{Burnett95,Burnett91}. Therefore, it would
seem that their maximal development should contain a maximal
Cauchy surface. We show that it does.
{\it Theorem 1.} The maximal development of any spherically symmetric
spacetime with collisionless matter (obeying the Vlasov equation) or a
massless scalar field (obeying the massless wave equation) that
possesses a CMC $S^1 \times S^2$ Cauchy surface $\Sigma$ admits a
unique foliation by CMC Cauchy surfaces with the mean curvature taking
on all real values. In particular, it contains a maximal Cauchy
surface and its singularities are crushing singularities.
By the maximal development of a globally hyperbolic spacetime, we mean
the maximal development of an initial data set induced on a Cauchy
surface in the spacetime. This is well-defined as the maximal
developments associated with any two Cauchy surfaces are necessarily
isometric \cite{CBGeroch69}. Further, recall that a spacetime with
compact Cauchy surfaces is said to have a future (past) crushing
singularity if the spacetime can be foliated by Cauchy surfaces such
that the mean curvature of these surfaces tends to infinity (negative
infinity) uniformly to the future (past). That the future and past
singularities associated with the spacetimes of theorem~1 are crushing
is then a simple consequence of the existence of a CMC foliation
taking on all real values.
As a consequence of theorem~1, the maximal development of the
spacetimes studied is rather simple. They expand from an initial
crushing singularity to a maximal hypersurface and then recollapse to
a final crushing singularity---all in a finite time. That is, they
satisfy the closed-universe recollapse conjecture in its strongest
sense as well as the closed-universe foliation conjecture.
While the maximal development of the spacetimes in theorem~1 is about
as complete as one could expect given the existence of a complete CMC
foliation, these spacetimes may still be extendible (though there is
no globally hyperbolic extension). In other words, theorem~1 does not
eliminate the possibility that these spacetimes violate cosmic
censorship. In fact, cosmic censorship is violated in the vacuum
case. This is easily seen by realizing that the maximal development
in this case is either of the of the two regions where $r < 2M$ of an
extended Schwarzschild spacetime of mass $M$ ($r$ is the areal
radius), modified by identifications so that the Cauchy surface
topology is $S^1 \times S^2$. Although the ``singularity''
corresponding to $r \to 2M$ is a crushing singularity, this is
actually a Cauchy horizon. Is this vacuum case exceptional? It is
worth noting that if a crushing singularity corresponds to $r \to 0$,
then the singularity must in fact be a curvature singularity. This
follows easily from the fact that $R_{abcd} R^{abcd} \ge (4m/r^3)^2$,
for any spherically symmetric spacetime satisfying the
null-convergence condition, and the fact that the mass function $m$ is
bounded away from zero by a positive constant in our case
\cite{Burnett91}. If we could show that $r$ must go to zero
(uniformly) at the extremes of our foliation, then the spacetime would
indeed be inextendible, thereby satisfying the cosmic censorship
hypothesis. Establishing such a result appears to be difficult and
the vacuum case shows that such a result will not always hold (though
this case may be exceptional). Using a different approach, Rein has
shown that for an open set of initial data, there is a crushing
singularity in which $r \to 0$ uniformly, and which, therefore, is a
curvature singularity \cite{Rein}. While this is encouraging, the
extent to which the spacetimes of theorem~1 satisfy cosmic censorship
remains to be seen.
The proof of theorem~1 involves a combination of three ideas. First,
it is known that spherically symmetric spacetimes with $S^1 \times
S^2$ or $S^3$ Cauchy surfaces and satisfying the dominant energy and
non-negative pressures (or merely ``radial'' non-negative pressure)
conditions have finite lifetimes \cite{Burnett95,Burnett91}. Second,
using a general theorem (which is independent of symmetry assumptions)
established in Sec.~\ref{sec:volume}, it follows that the spatial
volumes of Cauchy surfaces in the spacetime are bounded above, which
allows us to bound various fields describing the spacetime geometry.
Third, introducing a new time function to avoid the problems
associated with ``degenerate'' maximal hypersurfaces (i.e., surfaces
where the mean curvature cannot be used as a good coordinate), the
theorem then follows using the methods developed in \cite{Rendall95}.
Furthermore, it is worth noting that our method uses only a few
properties of the matter fields themselves. Namely, we use the fact
that they satisfy the dominant energy and ``radial'' non-negative
pressures conditions and, roughly speaking, the fact that the matter
fields are nonsingular as long as the spacetime metric is nonsingular.
This latter property has not been given a precise formulation, as it
seems difficult to do so, and serves merely as a heuristic
principle---the arguments for collisionless matter and the massless
scalar field in \cite{Rendall95} providing an example of what it means
in practice.
In theorem~1 we have restricted ourselves to spacetimes with $S^1
\times S^2$ Cauchy surfaces and have not considered similar spacetimes
with $S^3$ Cauchy surfaces. The problem with the $S^3$ case is that
there exist two timelike curves on which the symmetry orbits
degenerate to points. When we then pass to the quotient of our
spacetime by the symmetry group, the field equations on the quotient
spacetime are singular on boundary points corresponding to the
degenerate orbits. Experience has shown that this degeneracy can have
nontrivial consequences on the evolution of the spacetime. For
example, in the study of the spherically symmetric asymptotically flat
solutions of the Einstein-Vlasov equations, it has been shown that if
a solution of these equations develops a singularity, then the first
singularity (as measured in a particular time coordinate) is at the
center \cite{RRS}. However, currently it is not known how to decide
when a central singularity must occur. In the case of asymptotically
flat spherically symmetric solutions of the Einstein equations coupled
to a massless scalar field, Christodoulou has shown that naked
singularities do form in the center of symmetry for certain initial
data (and that they can form nowhere else) \cite{Christodoulou}. Note
that the degeneracy of the orbits in these spacetimes is of the same
type that occurs in the spherically symmetric spacetimes with $S^3$
Cauchy surfaces. Similar problems occur in the study of the vacuum
spacetimes with $U(1) \times U(1)$ symmetry and having $S^3$ or $S^1
\times S^2$ Cauchy surfaces. Here the dimension of the orbits is
non-constant and, consequently, this case is much harder to analyze
than the $T^3$ case, which has orbits of constant dimension
\cite{Chrusciel}. The spherically symmetric spacetimes with $S^1
\times S^2$ Cauchy surfaces, having no degenerate orbits, avoid these
complications.
It would, of course, be preferable to strengthen theorem~1 by removing
the requirement that there exist a CMC Cauchy surface in the
spacetime. While such a result seems plausible, the methods currently
used are not adequate to cover this more general case. Strengthening
our results in this direction is a subject for future research.
Our conventions are those of \cite{Wald84}, with the notable exception
that trace $H$ of the extrinsic curvature $K_{ab}$ of a spatial
hypersurface measures the {\em convergence} of the hypersurface to the
future. Thus, surfaces with negative $H$ are expanding to the future,
while those with positive $H$ are contracting to the future.
\section{Proof of theorem~1} \label{sec:proof}
Fix a spacetime $(M,g)$ satisfying the conditions of theorem~1. Both
classes of spacetimes considered here (the Einstein-Vlasov and
massless scalar field spacetimes) satisfy the dominant energy
condition (the Einstein tensor $G_{ab}$ satisfies $G_{ab}v^aw^b\ge 0$
for all future-directed timelike vectors $v^a$ and $w^b$) as well as
the timelike-convergence condition (the Ricci tensor satisfies
$R_{ab}t^a t^b \ge 0$ for all timelike $t^a$). While the
Einstein-Vlasov spacetimes also satisfy the non-negative pressures
condition ($G_{ab}x^ax^b \ge 0$ for all spacelike $x^a$), in general
the massless scalar field spacetimes do not. However, they do satisfy
the weaker ``radial'' non-negative pressures condition ($G_{ab}x^a x^b
\ge 0$ for all spatial vectors $x^a$ perpendicular to the spheres of
symmetry). It was shown in \cite{Burnett95,Burnett91} that the
spherically symmetric spacetimes with $S^3$ or $S^1 \times S^2$ Cauchy
surfaces satisfying the dominant energy and the non-negative pressures
conditions (or merely the ``radial'' non-negative pressures condition)
have a finite lifetime, i.e., the supremum of the lengths of all causal
curves is finite. Therefore, our spacetime $(M,g)$ has a finite
lifetime. It then follows immediately from lemma~2 (established in
Sec.~\ref{sec:volume}) that the volumes of all spatial Cauchy surfaces
in $(M,g)$ are bounded above.
Denote the mean curvature of the Cauchy surface $\Sigma$ by $t_0$.
This initial data surface must be spherically symmetric. In the case
$t_0 \neq 0$, this follows from the uniqueness theorem for such
hypersurfaces (see, e.g., theorem~1 of \cite{MarsdenTipler80}) since
if a rotation did not leave $\Sigma$ invariant, we would have a
distinct CMC Cauchy surface with identical (nonzero) constant mean
curvature. The case where $t_0 = 0$ then follows from the fact that
there is a neighborhood $N$ of $\Sigma$ in $M$ such that $N$ can be
foliated by CMC hypersurfaces, each having a different CMC, and the
fact that those with non-zero CMC must be spherically symmetric. As
the theorem has already been proven in the case where $t_0 = 0$
($\Sigma$ is a maximal hypersurface) \cite{Rendall95}, we shall take
$t_0$ to be negative ($\Sigma$ is expanding to the future). The case
where the mean curvature is initially positive follows by a
time-reversed argument. As was shown in \cite{Rendall95}, in a
neighborhood of the hypersurface $\Sigma$, the spacetime can be
foliated by CMC Cauchy surfaces. Define the scalar field $t$ at any
point to be the value of the mean curvature of the CMC hypersurface
passing through that point, i.e., so level surfaces of $t$ are CMC
hypersurfaces and, in particular, the surface $t=t_0$ is $\Sigma$. A
further scalar field $x$ can then be introduced so that the spacetime
metric $g$ is given by
\begin{equation} \label{metric}
g = -\alpha^2 {\rm d} t^2 + A^2[({\rm d} x+\beta {\rm d} t)^2 + a^2
\Omega],
\end{equation}
where $\Omega$ is the natural unit-metric associated with the spheres
of symmetry. The functions $\alpha$, $\beta$, and $A$ depend only on
$t$ and $x$ (being spherically symmetric) and are periodic in $x$.
The function $a$ depends only on $t$. The fields can be chosen so
that $\int \beta(t,x)\; {\rm d} x = 0$ for each $t$, where the
integral is taken over one period of a surface of constant $t$.
It was shown in \cite{Rendall95} that the initial data induced on
$\Sigma$ can be evolved so that $t$ covers the interval $(-\infty, 0)$
and that, if it can be evolved to the closed interval $(-\infty,0]$,
i.e., a maximal hypersurface is attained, the spacetime can be
extended and foliated by CMC spatial hypersurfaces taking on all real
values. Therefore, our task is to establish the existence of a
maximal hypersurface. To this accomplish this, we establish the
existence of upper bounds on $a$, $A$, and their inverses on the
interval $[t_0,0)$. We then introduce a new time function $\tau = f
\circ t$ by introducing a function $f$ that allows us to avoid the
problem associated with $t$ being a bad coordinate on maximal
hypersurfaces. Once this has been accomplished, theorem~1 will follow
from an argument similar to that used in \cite{Rendall95}.
First, we establish upper bounds on the area radius $r = aA$, the mass
function $m = {1 \over 2} r(1-\nabla^a r\nabla_a r)$, the volume
$V(t)$ of level surfaces of $t$, and their inverses. That $r$ and
$m^{-1}$ are bounded above follows from the results of
\cite{Burnett91}. (Note, $m$ is positive.)\ \ Further, the technique
introduced in \cite{MalecOMurchadha94} was used in \cite{Rendall95} to
show that $m/r$ is bounded above on $[t_0,0)$. Therefore, $m$ and
$r^{-1}$ are also bounded above on $[t_0,0)$. (That is, the mass $m$
cannot become arbitrarily large and $r$ cannot become arbitrarily
small in this portion of the spacetime. This is nontrivial as both
$m$ and $r^{-1}$ can become arbitrarily large on unbounded intervals, e.g.,
near an initial or final singularity.)\ \ As we have already
established that the volume of all spatial Cauchy surfaces are bounded
above, $V(t)$ is bounded above. Using the fact that $\partial_t V(t)$
is positive on $[t_0,0)$, as these hypersurfaces are everywhere
expanding, shows that $V$ is bounded from below by a positive
constant, and hence $V^{-1}$ is bounded above on $[t_0,0)$.
Next, that $a$, $A$, and their inverses are bounded above on $[t_0,0)$
now follows easily from the facts that $r=aA$,
\begin{equation}
V(t) = 4\pi \int a^2 A^3 \;{\rm d} x =
4\pi a^{-1} \int r^3 \;{\rm d} x,
\end{equation}
and our upper bounds for $V$, $r$, and their inverses.
Next, we bound $\alpha'$ using the lapse equation
\begin{equation} \label{lapse}
- A^{-3} (A \alpha')' + ( K_{ab}K^{ab} + R_{ab} n^a n^b)\alpha = 1,
\end{equation}
where $K_{ab}$ is the extrinsic curvature of the CMC hypersurface,
$n^a$ is a unit timelike normal to the CMC hypersurface, and a prime
denotes a derivative by $\partial_x$. (This is equation~(2.4) in
\cite{Rendall95}.)\ \ Using the fact that $K_{ab}K^{ab}$ is manifestly
non-negative and $R_{ab}n^an^b \ge 0$ by the timelike convergence
condition, it follows that $(A\alpha')' \ge - A^3$. Using the fact
that $A$ is bounded above and integrating in a CMC hypersurface, we
find that $(A \alpha') |_p - (A \alpha') |_q \ge -C_1$ for some
positive constant $C_1$ and any two points $p$ and $q$ in the
hypersurface. Choosing $q$ where $\alpha$ is extremal on the surface
(so $\alpha'(q)=0$) and using the fact $A^{-1}$ is bounded above shows
that $\alpha'$ is bounded from below. Choosing $p$ where $\alpha$ is
extremal on the surface (so $\alpha'(p)=0$) and using the fact
$A^{-1}$ is bounded above shows that $\alpha'$ is bounded from above.
Therefore, there exists a constant $C_2$ such that $|\alpha'| \le
C_2$. Thus, even if $\alpha$ is unbounded, it must diverge in a way
that is uniform in space: For any two points $p$ and $q$ in a CMC
hypersurface, $|\alpha(p) - \alpha(q)| = |\int_p^q \alpha' \;{\rm d}
x| \le \int_p^q |\alpha'| \;d x \le \pi C_2$.
If we knew that $\alpha$ were bounded above on $[t_0,0)$, we could
then proceed to argue as in \cite{Rendall95}. While such a bound can
be established rather easily for fields satisfying the dominant energy
and non-negative pressures conditions, such an argument fails for the
massless scalar field. The difficulty in establishing an upper bound
on $\alpha$ is linked to the possibility that ${\rm d} t$ may be zero
on a maximal hypersurface, and thus $t$ being a bad coordinate. Note
that this can only occur if $K_{ab}=0$ everywhere on $\Sigma$ (i.e.,
$\Sigma$ is momentarily static) and $R_{ab}n^a n^b = 0$ everywhere on
$\Sigma$. If the non-negative energy condition ($G_{ab}t^at^b \ge 0$
for all timelike $t^a$) and non-negative sum-pressures condition
[$G_{ab}(t^at^b+g^{ab}) \ge 0$ for all unit-timelike $t^a$] are
satisfied, then $R_{ab}n^a n^b = 0$ implies that $G_{ab}n^an^b =0$
and, hence, by the Hamiltonian constraint equation, the Ricci scalar
curvature of the metric induced on $\Sigma$ must be zero. However, it
is easy to show that there are no such spherically symmetric
geometries on $S^1 \times S^2$. Thus, the Einstein-Vlasov spacetimes
do not admit such surfaces. However, it can be shown that there are
massless scalar field spacetimes with such ``degenerate'' maximal
hypersurfaces. To avoid this difficulty, we change our time function
to one that is guaranteed to be well-behaved even on a maximal
hypersurface with ${\rm d} t=0$.
Fix any inextendible timelike curve $\gamma$ that is everywhere
orthogonal to the CMC hypersurfaces. The length of the segment of
$\gamma$ between any two CMC hypersurfaces $t=t_1$ and $t=t_2$ is then
simply $\int_{t_1}^{t_2} \alpha(\gamma(u)) \;{\rm d} u$. Using the
fact that there is a finite upper bound on the lengths of all timelike
curves in our spacetime, the integral
\begin{equation}
\int_{t_1}^0 \alpha(\gamma(u)) \;{\rm d} u = \lim_{t_2 \to 0}
\int_{t_1}^{t_2} \alpha(\gamma(u)) \;{\rm d} u
\end{equation}
must exist, i.e., $\alpha(\gamma(t))$ is integrable on any interval of
the form $[t_1,0)$. Fix some value $x_0$ of $x$ and consider the
function $\alpha(t,x_0)$. Since $\alpha'$ is bounded there is a
constant $C$ such that $\alpha(t,x_0)\le\alpha(\gamma(t))+C$. It
follows that $\alpha(t,x_0)$ is also integrable on any interval of the
form $[t_1,0)$. Using this fact, define the function $f$ on
$(-\infty,0)$ by setting
\begin{equation} \label{diffeo}
f(\lambda) = \lambda - \int_\lambda^0 \alpha(u,x_0) \;{\rm d} u.
\end{equation}
Noting that $f'(\lambda) = 1 + \alpha(\lambda,x_0)$ and
$\lim_{\lambda \to 0}f(\lambda) = 0$, we see that $f$ is an
orientation-preserving diffeomorphism from $(-\infty,0)$ to
$(-\infty,0)$. Hence,
\begin{equation} \label{newtime}
\tau = f \circ t
\end{equation}
is a new time function on our spacetime. Note that
${\partial \tau / \partial t} = 1 + \alpha(t,x_0)$.
The level surfaces of $\tau$ clearly coincide with those of $t$ and so
are CMC hypersurfaces. As a consequence the field equations for the
geometry and the matter written in terms of $\tau$ look very similar
to those written in terms of $t$. Using $\tau$ in place of $t$, the
metric has the same form as before
\begin{equation} \label{metric2}
g = -\tilde{\alpha}^2 {\rm d} \tau^2 +
A^2[({\rm d} x+\tilde{\beta} {\rm d} \tau)^2 + a^2 \Omega],
\end{equation}
where the new lapse function $\tilde{\alpha}$ is given by
\begin{equation} \label{alphat}
\tilde{\alpha} = \alpha \left({\partial t \over \partial \tau}\right)
= {\alpha \over 1 + \alpha(t,x_0)},
\end{equation}
and similarly for the new shift $\tilde{\beta}$. In terms of our new
coordinates ($\tau$ replacing $t$) and new variables ($\tilde{\alpha}$
and $\tilde{\beta}$ replacing $\alpha$ and $\beta$, respectively), the
field equations are the same as in \cite{Rendall95} with
$\partial_\tau$ replacing $\partial_t$, $\tilde{\alpha}$ replacing
$\alpha$, $\tilde{\beta}$ replacing $\beta$, and $\partial
t/\partial\tau$ replacing the right-hand side of
equation~(\ref{lapse}). Explicit occurrences of $t$ in the equations
are left unchanged, $t$ being simply considered as a function of
$\tau$, determined implicitly by equation~(\ref{newtime}). Using
equation~(\ref{alphat}), it is straightforward to show that $\partial
t / \partial \tau = 1 - \tilde{\alpha}(\tau,x_0)$. With this, the
lapse equation can be written as
\begin{equation} \label{newlapse}
-A^{-3} (A \tilde\alpha')' + ( K_{ab}K^{ab} + R_{ab} n^a n^b)\tilde\alpha
= 1 - \tilde\alpha(\tau,x_0).
\end{equation}
Using the fact that $\alpha'$ is bounded, as argued above, it follows
that $\alpha (t,x)\le \alpha(t,x_0) + C$, where $C$ is a constant.
Therefore, by equation~(\ref{alphat}), $\tilde{\alpha}$ is bounded
above.
It is now possible to apply the same type of arguments to the system
corresponding to the time coordinate $\tau$ as were applied in
\cite{Rendall95} to the system corresponding to the time coordinate
$t$ to show that all the basic geometric and matter quantities in the
equations written with respect to $\tau$ are bounded and that the same
is true for their spatial derivatives of any order. Bounding time
derivatives of all these quantities requires some more effort. All
but one of the steps in the inductive argument used to bound time
derivatives in \cite{Rendall95} apply without change. (Note that in
\cite{Rendall95}, derivatives with respect to $t$ were bounded,
whereas here, derivatives with respect to $\tau$ are bounded.)\ \ The
argument that does not carry over is that which was used to bound time
derivatives of $\alpha$ and $\alpha'$. To see why, consider the
equation obtained by differentiating equation~(\ref{newlapse}) $k$
times with respect to $\tau$
\begin{eqnarray} \label{derivatives}
- A^{-3} (A (D^k_\tau \tilde\alpha)')'
+ ( K_{ab}K^{ab} + R_{ab} n^a n^b) D^k_\tau \tilde\alpha\nonumber\\
+ D^k_\tau \tilde\alpha(\tau,x_0) = B_k,
\end{eqnarray}
where $D^k_\tau = \partial^k_\tau$ denotes the $k$-th partial
derivative with respect to $\tau$. Here $B_k$ is an expression
which is already known to be bounded when we are at the step in the
inductive argument to bound $D^k_\tau\tilde\alpha$ and
$D^k_\tau\tilde\alpha'$. In lemma~3.4 of \cite{Rendall95}, $D^k_t
\alpha$ was bounded by using the fact that $t$ was bounded away from
zero. The analogous procedure is clearly not possible in the present
situation, where $t$ is tending to zero. This kind of argument was
also used in \cite{Rendall95} to bound time derivatives of higher
order spatial derivatives of $\alpha$, but that is unnecessary, since
such bounds can be obtained directly by differentiating the lapse
equation once the time derivatives of $\alpha$ and $\alpha'$ have been
bounded. The same argument applies here, so all we need to do is to
prove the boundedness of $D^k_\tau\tilde\alpha$ and
$D^k_\tau\tilde\alpha'$ using equation~(\ref{derivatives}) under the
hypothesis that $B_k$ is bounded. This follows by simply noting that
equation~(\ref{derivatives}) has the same form for each value of $k$
and the following lemma.
{\it Lemma 1.} Consider the differential equation
\begin{equation} \label{ODE}
(au')'=bu+c+du(x_0)
\end{equation}
where $a$, $b$, $c$, $d$, and $u$ are $2\pi$-periodic functions on the
real line and $x_0$ is a point therein. Suppose that $a>0$, $b\ge 0$,
$d\ge 0$, and that $d$ is not identically zero. Then $|u|$ and $|u'|$
are bounded by constants depending only on the quantities $K_1 =
\max\{a^{-1}(x)\} > 0$, $K_2=\int_0^{2\pi} |c(x)| \;{\rm d} x \ge 0$,
$K_3 = \int_0^{2\pi} d(x)\;{\rm d} x > 0$, and $K_4 = \int_0^{2\pi}
b(x) \;{\rm d} x \ge 0$.
{\it Proof.} First, if $u(x_0) > 2\pi K_1K_2$, then $u > 0$
everywhere. To see this, suppose otherwise and let $x_1$ be a point
where $u$ achieves its maximum, so $u(x_1) \ge u(x_0) > 2\pi K_1K_2$
and let $x_2$ be that number such that $u > 0$ on $[x_1,x_2)$ and
$u(x_2)=0$ (so $x_1 < x_2 < x_1 + 2\pi$). Then, on the interval
$[x_1,x_2]$, we have $(au')' \ge c$, from which it follows that $u'
\ge -K_1K_2$ on $[x_1,x_2]$. Integrating this and using the fact that
$u(x_2)=0$, we find that $u(x_1) \le 2\pi K_1K_2$, contradicting the
fact that $u(x_1) > 2\pi K_1K_2$. Therefore, as $u$ is everywhere
positive, it follows that $(au')'\ge c$. Integrating this inequality
starting (or ending) at a point where $u' = 0$ shows that $|u'| \le
K_1K_2$. Integrating equation~(\ref{ODE}) from $0$ to $2\pi$ and
using the fact that $u$ is positive gives $u(x_0)\int_0^{2\pi} d(x)
\;{\rm d} x \le \int_0^{2\pi} |c(x)| \;{\rm d} x$, and hence,
$|u(x_0)| \le K_2 K_3^{-1}$. Using this and the fact that $|u'| \le
K_1K_2$ shows that $|u| \le K_2 K_3^{-1} + 2\pi K_1K_2$. Second, if
$u(x_0) < -2\pi K_1K_2$, a similar argument shows that $u$ is
everywhere negative and we again obtain the same bounds on $|u'|$ and
$|u|$. Third, suppose that $|u(x_0)| \le 2\pi K_1K_2$. If $\max(u) >
2\pi K_1K_2(1 + 2\pi K_1K_3)$, using the inequality $(au')' \ge c +d
u(x_0)$, we can argue much as before to see that $u$ is everywhere
positive and again obtain the same bounds on $|u'|$ and $|u|$.
Similarly, if $\min(u) < - 2\pi K_1K_2(1 + 2\pi K_1K_3)$, it follows
that $u$ is everywhere negative and again we recover the same bounds
on $|u'|$ and $|u|$. Next, if $|u|
\le 2\pi K_1K_2(1 + 2\pi K_1K_3)$ everywhere, $|u|$ is already
bounded, and to bound $|u'|$, we note that we have bounds for all
terms on the right hand side of equation~(\ref{ODE}), so it suffices
to integrate it starting from a point where $u'$ is zero to bound
$|u'|$.$\Box$
At this stage, we have indicated how all geometric and matter
quantities, expressed in terms of the new time coordinate $\tau$,
can be bounded, together with all their derivatives. In particular,
this means that all these quantities are uniformly continuous
on any interval of the form $[\tau_1,0)$, where $\tau_1$
is finite. It follows that all these quantities have smooth extensions
to the interval $[\tau_1,0]$. Restricting them to the hypersurface
$\tau=0$ gives a initial data set for the Einstein-matter
equations with zero mean curvature. By the standard uniqueness
theorems for the Cauchy problem, the spacetime which, in the old
coordinates, was defined on the interval $(-\infty,0)$ is isometric
to a subset of the maximal development of this new initial data
set. It follows that the original spacetime has an extension which
contains a maximal hypersurface.
Lastly, that the foliation is unique now follows from the fact that
compact CMC Cauchy surfaces with non-zero mean curvature are unique
\cite{MarsdenTipler80} and that the spacetime is indeed maximal
follows from the fact that any spacetime admitting a complete foliation
by compact CMC Cauchy surfaces is maximal \cite{EardleySmarr79}.
\section{A bound for the volume of space} \label{sec:volume}
It is well known that as we transport an ``infinitesimal'' spacelike
surface $S$ along the geodesics normal to itself, the ratio $\nu$ of
its volume of to its original volume is governed by the Raychaudhuri
equation
\begin{equation} \label{ray}
{{\rm d}^2 \over {\rm d} t^2} \nu^{1/3} + {1 \over 3}
\left( R_{ab}t^at^b + \sigma_{ab}\sigma^{ab} \right) \nu^{1/3} = 0,
\end{equation}
where $t$ is the proper time measured along the geodesics normal to
$S$, $R_{ab}$ is the Ricci tensor, and $\sigma_{ab}$ is the shear
tensor associated with the geodesic flow
\cite{HawkingEllis73,Wald84,Penrose72}. (This equation is usually
written in terms of the divergence of the geodesic flow $\theta =
\nu^{-1} d\nu/dt$.)\ \ On the surface $S$, $\nu$ satisfies the initial
condition $\nu = 1$ and $d\nu/dt = - H(p)$, where $H(p)$ is the trace
of the extrinsic curvature of $S$ at the point $p$ where the geodesic
intersects $S$. Therefore, if the spacetime satisfies the
timelike-convergence condition ($R_{ab}t^at^b \ge 0$ for all timelike
$t^a$), it follows that as long as $\nu$ remains non-negative,
\begin{equation}
{{\rm d}^2 \over {\rm d} t^2} \nu^{1/3} \le 0,
\end{equation}
from which we find that
\begin{equation} \label{localbound}
\nu(t) \le \left[1 - {1 \over 3} H(p) (t-t_0) \right]^3.
\end{equation}
This equation bounds the growth of the volume of a local spatial
region in the spacetime.
Using this result, it is not difficult to show that, in a spacetime
satisfying the timelike-convergence condition, if we fix a Cauchy
surface $\Sigma_0$ and construct from it a second Cauchy surface
$\Sigma$ by transporting $\Sigma_0$ to the future along the flow
determined by the geodesics normal to $\Sigma_0$, as long as these
flow lines do not self-intersect (which will be true if $\Sigma$ is
sufficiently close to $\Sigma_0$), then
\begin{equation} \label{specialglobalbound}
\text{vol}(\Sigma) \le \text{vol}(\Sigma_0) \left[ 1 + {1 \over 3}
\sup_{\Sigma_0}(-H) T \right]^3,
\end{equation}
where $\text{vol}(S)$ denotes the three-volume of a Cauchy surface $S$
and $T$ is the ``distance'' between the two surfaces measured by the
lengths of the geodesics normal to $\Sigma_0$ (which will be
independent of which geodesic is chosen by the construction of
$\Sigma$). Therefore, we have a bound on the volume of $\Sigma$ in
terms of the volume of $\Sigma_0$, the extrinsic curvature of
$\Sigma_0$, and the distance between $\Sigma_0$ and $\Sigma$. Does a
similar result hold for more general Cauchy surfaces $\Sigma$? For
instance, a more general hypersurface $\Sigma$ may not be everywhere
normal to the geodesics from $\Sigma_0$, some geodesics normal to
$\Sigma_0$ may intersect one another between $\Sigma_0$ and $\Sigma$,
and parts of $\Sigma$ may lie to the future of $\Sigma_0$ while other
parts may lie to the past. Can the simple bound given by
equation~(\ref{specialglobalbound}) be modified to cover these
cases? That it can is the subject of the following lemma.
{\it Lemma 2.} Fix an orientable globally hyperbolic spacetime
$(M,g_{ab})$ satisfying the timelike-convergence condition
($R_{ab}t^at^b \ge 0$ for all timelike $t^a$) and a smooth spacelike
Cauchy surface $\Sigma_0$ therein. Then, for any smooth spacelike
Cauchy surface $\Sigma$,
\begin{equation} \label{boundany}
\text{vol}(\Sigma) \le \text{vol}(\Sigma_0)
\left[ 1 + {1 \over 3} \sup_{\Sigma_0}(|H|)
\Delta(\Sigma_0,\Sigma) \right]^3,
\end{equation}
where $\text{vol}(S)$ denotes the three-volume of a Cauchy surface
$S$, $H$ is the trace of the extrinsic curvature of $\Sigma_0$ (using
the convention that $H$ measures the {\it convergence} of the {\it
future-directed} timelike normals to a spacelike surface), and
$\Delta(\Sigma_0,\Sigma)$ is the least upper bound to the lengths of
causal curves connecting $\Sigma_0$ to $\Sigma$ (either future or past
directed). Further, for any Cauchy surface $\Sigma \subset
D^+(\Sigma_0)$,
\begin{equation} \label{boundfuture}
\text{vol}(\Sigma) \le \text{vol}(\Sigma_0)
\left[ 1 + {1 \over 3} \sup_{\Sigma_0}(-H)
\Delta(\Sigma_0,\Sigma) \right]^3.
\end{equation}
Note that for $p,q \in M$, $\Delta(p,q)$ is not quite the distance
function $d(p,q)$ as used in \cite{HawkingEllis73} as $d(p,q) =
0$ if $q \in J^-(p)$. Instead, $\Delta(p,q)$ does not distinguish
between future and past: $\Delta(p,q) = \Delta(q,p) = d(p,q) +
d(q,p)$.
{}From lemma~2, we see that for a spacetime satisfying the
timelike-convergence condition, possessing compact Cauchy surfaces,
and having a finite lifetime (in the sense that $d(p,q)$ [equivalently
$\Delta(p,q)$] is bounded above by a constant independent of $p$ and
$q$), then the volume of a Cauchy surface therein cannot be
arbitrarily large. Further, we see that if the spacetime admits a
maximal Cauchy surface $\Sigma_0$ ($H=0$ thereon), we reproduce the
result that there is no other Cauchy surface having volume larger than
$\Sigma_0$ (though there may be surfaces of equal volume)
\cite{MarsdenTipler80}.
In the following, ${\rm d} f$ denotes the derivative map associated
with a differentiable map $f$ between manifolds. When viewed as a
pull-back, we denote ${\rm d} f$ by $f^*$ and, when viewed as a
push-forward, we denote ${\rm d} f$ by $f_*$. For a map $f: A \to B$,
$f[A]$ denotes the image of $A$ in $B$. Lastly, $A \setminus B$
denotes the set of elements in $A$ that are not in $B$.
\subsection{Proof of lemma 2} \label{sec:prooflemma2}
To begin the proof of lemma~2, for each point $p \in \Sigma_0$, let
$\gamma_p$ denote the unique inextendible geodesic containing $p$ and
intersecting $\Sigma_0$ orthogonally. Parameterize $\gamma_p$ by $t$
so that the tangent vector to $\gamma_p$ is future-directed
unit-timelike and $\gamma_p(0)=p$. Then, define the map $f:\Sigma_0
\to \Sigma$, by
\begin{equation}
f(p) = \gamma_p \cap \Sigma.
\end{equation}
Note that for each $p \in \Sigma_0$, $f$ is well defined since
$\gamma_p$ intersects $\Sigma$ at precisely one point as $\Sigma$
is a spacelike Cauchy surface for the spacetime.
Next, let ${\cal K}$ be the subset of $\Sigma_0$ defined by the
property that $p \in {\cal K}$ if and only if the geodesic $\gamma_p$
does not possess a point conjugate to $\Sigma_0$ between $\Sigma_0$
and $\Sigma$ (although it may have such a conjugate point on
$\Sigma$). Note that this is precisely the condition that for each $p
\in {\cal K}$ the solution $\nu$ to equation~(\ref{ray}) along
$\gamma_p$, satisfying the initial conditions $\nu = 1$ and $d\nu/dt =
H(p)$ at $p$, be strictly positive on the portion of $\gamma_p$
between $p$ and $f(p)$. It follows that ${\cal K}$ is closed.
Furthermore, $f$ maps ${\cal K}$ onto $\Sigma$. To see this, recall
that for any point $q \in \Sigma$ there exists a timelike curve $\mu$
connecting $q$ to $\Sigma_0$ having a length no less than any other
such curve. Furthermore, such a curve $\mu$ must intersect $\Sigma_0$
normally, is geodetic, and has no point conjugate to $\Sigma_0$
between $\Sigma_0$ and $q$. (See Theorem~9.3.5 of \cite{Wald84}.)\ \
Therefore, the point $p = \mu \cap \Sigma_0$ is in ${\cal K}$ and $\mu
\subset \gamma_p$, so $f(p) = \gamma_p \cap \Sigma = \mu \cap \Sigma =
q$. Therefore, $f$ maps ${\cal K}$ onto $\Sigma$. However, in
general, $f$ will not be one-to-one between ${\cal K}$ and $\Sigma$.
Let $C$ denote the set of critical points of the map $f$ on
$\Sigma_0$. That is, $p \in C$ if and only if its derivative map
$f_*: (T\Sigma_0)_p \to (T\Sigma)_{f(p)}$ is not onto. Then, by
Sard's theorem \cite{Milnor65}, $f[C]$ (the critical values of $f$),
and hence $f[{\cal K} \cap C]$, are sets of measure zero on $\Sigma$.
Now, note that $\Sigma$ can be expressed as the union of $f[{\cal K}
\setminus C]$ and a set having measure zero. To see this, we write
\begin{eqnarray}
\Sigma & = & f[{\cal K}]
= f[({\cal K} \setminus C) \cup ({\cal K} \cap C)] \nonumber\\
& = & f[{\cal K} \setminus C] \cup
\left( f[{\cal K} \cap C] \setminus f[{\cal K} \setminus C] \right).
\end{eqnarray}
The last two sets are manifestly disjoint and the latter is a set of
measure zero (as it is a subset of a set of measure zero). Therefore, we
need only concern ourselves the behavior of $f$ on the set of regular
points of $f$ within ${\cal K}$. This is useful since, by the inverse
function theorem \cite{Milnor65}, $f$ is a local diffeomorphism
between ${\cal K} \setminus C$ and $f[{\cal K} \setminus C]$. As we
shall see, for all $p \in {\cal K}\setminus C$, the point $f(p)$ is
not conjugate to $\Sigma_0$ on $\gamma_p$, from which it follows that
${\cal K} \setminus C$ is an open subset of $\Sigma_0$.
Denote volume elements associated with the induced metrics on
$\Sigma_0$ and $\Sigma$ by $e_{abc}$ and $\epsilon_{abc}$,
respectively, chosen so that $e_{abc}$ and $\epsilon_{abc}$ correspond
to the same spatial orientation class (which can be done as the
spacetime is both time-orientable and orientable). Then the Jacobian
of the map $f$ is that unique scalar field $J$ on $\Sigma_0$ such that
\begin{equation} \label{defJ}
(f^*\epsilon)_{abc} = J e_{abc}.
\end{equation}
Note that $J$ is zero on $C$ and positive on ${\cal K} \setminus C$.
With these definitions, we have
\begin{eqnarray} \label{boundvol}
\text{vol}(\Sigma)
& = & \int_{f[{\cal K} \setminus C]} \epsilon \nonumber \\
& \le & \int_{{\cal K} \setminus C} (f^*\epsilon) \nonumber \\
& \le & \left[\sup_{{\cal K} \setminus C}(J)\right]
\int_{{\cal K} \setminus C} e \nonumber \\
& \le & \left[\sup_{{\cal K} \setminus C}(J)\right]
\text{vol}(\Sigma_0).
\end{eqnarray}
The first step follows from the facts that $\Sigma = f[{\cal K}]$ and
$f[{\cal K} \cap C]$ is a set of measure zero. That we have an
inequality in the second step follows from the fact that although $f$
is a local diffeomorphism, it may not be one-to-one between ${\cal K}
\setminus C$ and $f[{\cal K} \setminus C]$. The third step follows
from the definition of $J$ given by equation~(\ref{defJ}) and the fact
that $J$ is bounded above by its supremum. Lastly, the fourth step
follows from the fact that ${\cal K} \setminus C$ is a subset of
$\Sigma_0$. So, to prove lemma~2, we need to show that, on the set
${\cal K} \setminus C$, $J$ is bounded above by the relevant
expressions in lemma~2.
To that end, define $\phi: \Sigma_0 \times {\Bbb R} \to M$ by setting
$\phi(p,t) = \gamma_p(t)$. Of course, if $\gamma_p$ is not future and
past complete, this will not be defined for all $t$. Next, define
$T:\Sigma_0 \to {\Bbb R}$ by setting $T(p)$ to that number such that
$\gamma_p(T(p)) = f(p)$, i.e., $T(p)$ is the ``time'' along the
geodesic $\gamma_p$ at which $\gamma_p$ intersects $\Sigma$. Note
that if $f(p)$ lies to the future of $\Sigma_0$, then $T(p)$ is
positive, while if $f(p)$ lies to the past of $\Sigma_0$, then $T(p)$
is negative.
Fix a point $p \in {\cal K} \setminus C$ and define the map $g:
\Sigma_0 \to M$ by setting $g(q) = \phi(q,T(p))$. Should
$\gamma_q(T(p))$ not be defined, then $g$ is not defined for that
point of $\Sigma_0$. However, it will always be defined for some
neighborhood of $p$ as $g(p) = f(p)$. Notice that $g$ simply
``translates'' points on $\Sigma_0$ along the geodesics normal to
$\Sigma_0$ a fixed distance $T(p)$ (independent of point), i.e., it is
a translation along the normal geodesic ``flow''. Therefore, the
derivative map of $g$ at a point is precisely the geodesic deviation
map. In particular, ${\rm d} g$ is injective (one-to-one) from
$(T\Sigma_0)_p$ to $(TM)_{f(p)}$ if and only if $f(p)$ is not
conjugate to $\Sigma_0$ on $\gamma_p$ (by the definition of such a
conjugate point).
Noting that $f$ can be written as $f(q) = \phi(q,T(q))$, we see that
the derivative maps of $f$ and $g$ at $p$ [both of which are maps from
$(T\Sigma_0)_p$ to $(TM)_{f(p)}$] are related by
\begin{equation} \label{dmaps}
({\rm d} f)^a{}_b = ({\rm d} g)^a{}_b + t^a ({\rm d} T)_b,
\end{equation}
where $t^a$ is the unit future-directed tangent vector to $\gamma_p$
at $f(p)$. From this we see that ${\rm d} f$ is injective [from
$(T\Sigma_0)_p$ to $(TM)_{f(p)}$] if and only if ${\rm d} g$ is
injective. Therefore, on ${\cal K} \setminus C$, not only is ${\rm d}
f$ injective, but ${\rm d} g$ is also injective, and hence $f(p)$ is
not conjugate to $\Sigma_0$ on $\gamma_p$.
Define $\hat{e}_{abc}$ at $f(p)$ by parallel transporting $e_{abc}$ at
$p$ along $\gamma_p$. Then,
\begin{equation} \label{f^*e-hat}
(f^*\hat{e})_{abc} = (g^*\hat{e})_{abc} = \nu(T(p)) e_{abc}.
\end{equation}
The first equality follows from (\ref{dmaps}) and the fact that $t^a
\hat{e}_{abc} = 0$. The second equality follows by recognizing that
the coefficient of the right-hand most term is precisely the ratio of
the volume of an ``infinitesimal'' region in $\Sigma_0$ to its
original volume as it is transported along the geodesic flow normal to
$\Sigma_0$. As the transport is done from $p$ to $f(p)$, the
coefficient is $\nu(T(p))$, where $\nu$ is the solution of
equation~(\ref{ray}) satisfying the stated initial conditions. (In
other words, $\nu(t)$ is the Jacobian of the geodesic deviation map.)
Denote the future-directed normal to $\Sigma$ at $f(p)$ by $n^a$.
Then, there exists a unit-spacelike vector $x^a \in (T\Sigma)_{f(p)}$
such that $t^a = \gamma (n^a + \beta x^a)$, where $\gamma = (-t^a
n_a)$ and $\beta = \sqrt{1 - \gamma^{-2}}$. Then, for one of the two
volume elements $\epsilon_{abcd}$ on $M$ associated with the spacetime
metric, we have $\epsilon_{abc} = n^m \epsilon_{mabc}$ and
$\hat{e}_{abc} = t^m \epsilon_{mabc}$, which gives the following
relation between these two tensors at $f(p)$,
\begin{equation} \label{relate}
\hat{e}_{abc} = \gamma \epsilon_{abc} + \gamma\beta x^m\epsilon_{mabc}.
\end{equation}
Therefore,
\begin{equation} \label{relatepullbacke-hatepsilon}
(f^* \hat{e})_{abc} = \gamma (f^*\epsilon)_{abc},
\end{equation}
where we have used (\ref{relate}) and the fact that the pull-back of
$x^m \epsilon_{mabc}$ by $f$ must be zero as $x^m$ is in the surface
$\Sigma$ and the contraction of $\epsilon_{abcd}$ with four vectors
all in a three-dimensional subspace must be zero. Therefore, using
(\ref{relatepullbacke-hatepsilon}) and~(\ref{f^*e-hat}), we see that
\begin{equation} \label{f^*epsilon}
(f^*\epsilon)_{abc} = (-t^an_a)^{-1}\nu(T(p)) e_{abc},
\end{equation}
which when compared to (\ref{defJ}), gives
\begin{equation}
J(p) = (-t^an_a)^{-1}\nu(T(p)).
\end{equation}
Since $(-t^a n_a)^{-1} \le 1$ and $\nu(T(p))$ is bounded above by
(\ref{localbound}), we have
\begin{equation}
J(p) \le \left[ 1 - {1 \over 3} H(p)T(p) \right]^3.
\end{equation}
So, if $\Sigma \subset D^+(\Sigma_0)$, we have $0 \le T(p) \le
\Delta(\Sigma_0,\Sigma)$ and $-H(p) \le \sup_{\Sigma_0}(-H)$, and
therefore,
\begin{equation}
\sup_{{\cal K} \setminus C}(J) \le \left[1 + {1 \over 3}
\sup_{\Sigma_0}(-H)\Delta(\Sigma_0,\Sigma) \right]^3,
\end{equation}
which with (\ref{boundvol}) establishes
equation~(\ref{boundfuture}). More generally, as
\begin{equation}
H(p)T(p) \le |H(p)||T(p)| \le \sup_{\Sigma_0}(|H|)
\Delta(\Sigma_0,\Sigma),
\end{equation}
we have
\begin{equation}
\sup_{{\cal K} \setminus C}(J) \le \left[1 + {1 \over 3}
\sup_{\Sigma_0}(|H|)\Delta(\Sigma_0,\Sigma) \right]^3,
\end{equation}
which with (\ref{boundvol}) establishes equation~(\ref{boundany}).
This completes the proof of lemma~2.
|
\section{Introduction}
Transfer based approaches to machine translation (MT) involve three main
phases: analysis, transfer and generation. During analysis, the syntactic and
semantic structure of a sentence is made explicit through a source language
(SL) grammar and semantic processing modules. The result of analysis is one or
more syntactic and semantic representations which are used to construct a
syntactic and/or semantic representation in the target language (TL) through a
series of transfer rules and a bilingual lexicon. From this representation a TL
sentence is generated based on some form of mapping procedure, usually
exploiting the TL grammar \footnote{
While this definition of transfer systems is current in most MT discussions, it
has been challenged \acite{kayetal94} on the basis that the interlingua-transfer
distinction, that is, the distinction between systems which construct language
independent representations and systems which do not, is artificial and that in
fact the two paradigms simply represent different aspects of the same problem.
While we agree with this observation, many systems at present start with
an interlingua or a transfer architecture and then incorporate solutions from
the alternative paradigm. We therefore maintain the distinction, at least for
the purposes of this paper.}.
In this paper we describe a prototype implementation of a transfer MT system
based on the lexicalist MT (LMT) approach of \cite{whitelock92}, also
known as `Shake-and-Bake' (SB). For our implementation we have extended the
original SB formulation by postulating bilingual lexical rules (bi-lexical rules
henceforth)
which dynamically expand the bilingual lexicon in order to
extend its functionality. This allows us to uniformly treat mono- and
multi-lexeme translations in a variety of contexts.
We describe the main characteristics of the LMT approach. This is followed by a
description of the problems posed by certain multi-lexeme translations, and of
how bi-lexical rules, in conjunction with lexical semantic information provide
a framework for overcoming these problems. We then point out some limitations
in our approach and give some idea as to the status of our implementation.
\section{Lexicalist Machine Translation}
In its original formulation, LMT consists of three main phrases: analysis,
lexical-semantic transfer and generation. The analysis phase involves parsing
the input sentence to produce an output bag or multiset of SL lexical signs
instantiated with sufficient information to permit appropriate translation.
Transfer maps these signs into a TL bag through the bilingual lexicon in which
sets of source and target lexical signs are placed in translation
correspondence. Generation consists of finding an ordering of the TL bag
which satisfies the constraints imposed by the TL grammar. Normally,
generation involves a modified parser which ignores ordering information
\acite{brew92,popowich95} although other approaches are also possible
\acite{poznanskietal95}.
\subsection{Notation}
We introduce some notation through a simple example of our implementation.
Since we will not be concerned with quantification nor scoping, we adopt a
simplified transfer representation. If quantification and scope were to be
included, however, a mechanism along the lines of \cite{franketal95} and
\cite{copestakeetal95b} may be followed in order to preserve the recursiveless
nature of lexicalist transfer.
Our lexical signs broadly follow the signs of \cite{pollardetal87} although our
work seems adaptable to the signs of \cite{pollardetal94}. The implementation
is based on the Typed Features Structures (TFSs) of the Acquilex LKB
\acite{copestakeetal93} from where we borrow our notation. Consider the (simplified)
lexical entry for `John':
\begin{quote}
{\scriptsize
$\avmplus{\att{proper-name}\\
\attvaltyp{orth}{John}\\
\attval{syn}{\avmplus{\att{syn}\\
\attvaltyp{agr}{3sg}}}\\
\attvalshrunktyp{qualia}{qualia}\\
\attvaltyp{sem}{john1(x)}}$}
\end{quote}
In this TFS, features are written in small capitals, while types are in bold
face. To make TFSs easier to read, detail may be hidden by `shrinking' a TFS;
this is indicated with a box around the type of the TFS (e.g.
$\boxvaluebold{qualia}$ above). TFSs of type {\bf qualia} encode lexical
semantic information based on the Qualia structures of \cite{pustejovsky91}.
For the semantic representation of proper names we assume a predicate treatment
following the arguments of \cite[225]{devlin91}. A bilexical entry for `John
-- {\em Juan}' would be:
\begin{quote}
{\scriptsize
$\avmplus{\att{proper-name}\\
\attvaltyp{orth}{John}\\
\attval{syn}{\avmplus{\att{syn}\\
\attvaltyp{agr}{3sg}}}\\
\attvalshrunktyp{qualia}{qualia}\\
\attvaltyp{lang}{english}\\
\attvaltyp{sem}{john1(x)}}
\leftrightarrow
\avmplus{\att{proper-name}\\
\attvaltyp{orth}{Juan}\\
\attval{syn}{\avmplus{\att{syn}\\
\attvaltyp{agr}{3sg}}}\\
\attvalshrunktyp{qualia}{qualia}\\
\attvaltyp{lang}{spanish}\\
\attvaltyp{sem}{juan1(x)}}$}
\end{quote}
For reasons of space and convenience, we will abbreviate the above lexical sign
and bilexical entry to
\begin{exquote}
john1$_{x}$\\
john1$_{x}$ $\leftrightarrow$ juan1$_{x}$\\
\end{exquote}
respectively, where the subscripts correspond to the argument variable. It
should be emphasised, however, that this abbreviated notation implicitly
includes syntactic and semantic information which may be accessed during
transfer or generation.
To exemplify LMT, consider the translation of `John likes Mary'. Analysis
results in a list\footnote{We use lists of SL lexical items, instead of bags as
is done in SB, to avoid certain inefficiencies caused by the nature of
lexicalist transfer \acitec[221]{gareyetal79}.} of lexical signs the semantics
of which will contain shared variables:
\begin{exquote}
john1$_{x}$ love1$_{e,x,y}$ mary1$_{x}$
\end{exquote}
The (tenseless) FOL formula corresponding to this expression is $\exists exy$.
john1($x$) \& love1($e,x,y$) \& mary1($y$), but since quantification and scope
will be ignored they will be omitted from our examples; furthermore,
coordination will be assumed between predicates unless otherwise stated.
Before transfer, a process similar to skolemization is applied to the transfer
representation in order to replace variables by constants. The purpose of this
operation is to prevent spurious bindings during lexicalist generation, as will
become clearer later. The result of analysis is a list of lexical signs with
translationally relevant relationships expressed by shared constants (indicated
by integers in our notation):
\begin{exquote}
john1$_{1}$ love1$_{2,1,3}$ mary1$_{3}$
\end{exquote}
The transfer step uses the source side of the bilexicon (possibly expanded by
bilingual lexical rules as described below) to derive a total cover of the SL
list \acitec[221]{gareyetal79} (a total cover is a division of a set into a
number of allowed subsets such that every element in the set is a member of
exactly one subset; we extend the term here to apply it to lists). The
bilexicon below enables construction of an appropriate TL bag:
\begin{quote}
\footnotesize
john1$_{x}$ $\leftrightarrow$ juan1$_{x}$\\
mary1$_{x}$ $\leftrightarrow$ mar\'\i a1$_{x}$\\
love1$_{x,y,z}$ $\leftrightarrow$ amar1$_{x,y,z}$ a1$_{z}$
\end{quote}
(Tense is omitted in this example; a simplistic model has been adopted in which an
interlingua tense feature is passed from source to target verbs in the
bilexicon.) Note that we include function words such as the Spanish case
marker {\em a} in the bilingual lexicon (and therefore in the transfer
representation). These words are treated as vacuous predicates
\acite{calderetal89} over the variable of the semantic head on which they
depend. For the present example, transfer results in the following TL bag:
\begin{quote}
\footnotesize
\{juan1$_{1}$ , amar1$_{2,1,3}$ , a$_{3}$ , mar\'\i a1$_{3}$\}
\end{quote}
Lexicalist generation involves reordering the TL bag to construct a valid TL
sentence. Since normally all permutations of the TL bag are attempted, the fact
that variables are replaced by constants ensures that arguments not shared
between predicates in the SL representation are not shared in the TL
representation either. This prevents {\em Mar\'\i a} from being the subject of
the sentence. The result of generation, after morphological synthesis, is:
\begin{quote}
\footnotesize
{\em Juan ama a Mar\'\i a}
\end{quote}
\subsection{Other Properties of LMT}
LMT encourages two useful properties: modularity and reversibility. From an
engineering point of view, modularity is desirable because it can reduce
development and maintenance costs. By using sets of lexical signs as their
transfer representation, LMT systems can reduce the difficulties posed by
structural mismatches between two languages, thus increasing the independence
between source and target transfer representations. For example, transfer
systems adopting a recursive representation for transfer \acite{kaplanetal89},
as opposed to a non-recursive one \acite{copestakeetal95b}, may need additional
mechanisms for handling head switching \acite{kaplanetal93}. By contrast,
under a lexicalist approach, head switching can be handled purely
compositionally with minimal assumptions \acite{whitelock92}.
Reversibility is an important property in bi-directional systems as it reduces
development costs. In LMT, grammars are fully reversible since they are used in
similar ways for analysis and generation: the difference is that during
lexicalist generation, ordering information is disregarded. However, the
process is complete because the generator is guaranteed to generate
all the strings accepted by the TL grammar which satisfy the constraints
imposed by the TL bag. Lexicalist generation is also sound because only
strings which satisfy the constraints of the TL grammar are constructed. In
addition, termination is guaranteed if it is guaranteed for parsing since one
can at worst construct a generation algorithm which simply attempts all
permutations of the TL bag and then parses them in order to test whether they
are appropriate TL sentences.
\section{Multi-Lexical Translations}
\label{mul-tra-sec}
One of the reasons for transfer modules being expensive to construct is the
presence of complex transfer relations \acite{arnoldetal92,hutchinsetal92}. One
type of phenomena that leads to complex transfer in a number of systems may be
called multi-lexical translation. These are translations in which a phrase
cannot easily be translated through the translation of its parts. The
translation of idioms is an extreme case of this. For example, `kick the
bucket' translates as {\em estirar la pata} (Lit. `to stretch a leg') in
Spanish, even though there is no simple correspondence between the components
of each phrase (all translations in this paper are between English and Spanish
unless otherwise stated). For such constructions, structures corresponding to
the source and target phrases need to be equated either in the transfer module
\acite{schenk86} or in separate dictionaries \acite{sadleretal90} in many
systems. Other phenomena which may be loosely labelled multi-lexeme
translations include: lexical gaps such as `piece of advice' -- {\em consejo}
\acite{soleretal93}; support verb and category differences such as `to be
thirsty' -- {\em tener sed} (to have thirst) \acite{danlosetal92};
lexicalization patterns like `swim across the river Dee' -- {\em cruzar el
r\'\i o Dee nadando} \acite{talmy85}; conflational divergences as in `to stab
someone' -- {\em darle pu\~{n}aladas a alguien} \acite{dorr92}.
Phenomena such as idioms, lexical gaps and conflational divergences can be
tackled in LMT by equating sets of source and target lexical signs:
\begin{exquote}
a) kick1$_{e,s,o}$, the1$_{o}$, bucket1$_{o}$ $\leftrightarrow$ estirar1$_{e,s,o}$, la1$_{o}$, pata1$_{o}$\\
b) piece1$_{x}$, of1$_{x,y}$, advice1$_{y}$ $\leftrightarrow$ consejo1$_{x}$\\
c) stab1$_{e,s,o}$ $\leftrightarrow$ dar1$_{e,s,p,o}$ le1$_{o}$ pu\~nalada1$_{p}$ a1$_{o}$
\end{exquote}
(We include lexical signs for determiners, clitics and accusative markers as
predicates over the variable of their syntactic head; however, reasoning
formalisms may dispense with them.) Note that we choose the variable of
`piece' on the English side as the argument variable on the Spanish side; if
phrases such as `a piece of good advice' are allowed, the Spanish side would be
{\em consejo}1$_{x\sqcup y}$, whose semantic argument would be unifiable with
both $x$ and $y$ to permit modifiers and heads to combine appropriately during
generation.
To translate `John kicked the bucket', the SL transfer representation:
\begin{exquote}
john1$_{1}$ kick1$_{2,1,3}$ the1$_{3}$ bucket1$_{3}$
\end{exquote}
is covered by the bilexicon. The result is the union of the target side of all
the bilexical entries used in this process:
\begin{exquote}
\{juan1$_{1}$\} $\cup$ \{estirar1$_{2,1,3}$, la1$_{3}$, pata1$_{3}$\}
\end{exquote}
(We ignore the literal translation of the idiom.) Generation then proceeds via
the Spanish grammar and bag generator.
In the case of the other multi-lexeme translations mentioned the difficulties
posed by varying lexical elements in part or all of the translation relation
cannot be easily handled in the original SB formulation. Consider for example
the case of `John is thirsty'; its Spanish translation, {\em Juan tiene sed}
(lit. `John has thirst') differs from it in two main ways: the English
adjective translates into a Spanish noun, while the verb is not intuitively
felt to be the translation of {\em tener}. The problem for LMT based on
one-to-one transfer is that a literal translation into Spanish is incorrect
(*{\em Juan est\'a sediento}), and that even if TL filtering
\acite{alshawietal92} were used to eliminate such a sentence, the efficiency of
the system would be compromised and translation of unseen sentences would be
more error prone. Alternatively, an idiom-based translation in which the
bilexicon relates `be thirsty' and {\em tener sed} ignores important systematic
differences between the two languages:
\begin{exquote}
\begin{tabular}{ll}
John {\bf is thirsty} & Juan {\bf tiene sed}\\
John {\bf is hungry} & Juan {\bf tiene hambre}\\
John {\bf is lucky} & Juan {\bf tiene suerte}\\
John {\bf is angry} & Juan {\bf tiene rabia}\\
John {\bf is hot} & Juan {\bf tiene calor}\\
John {\bf is cold} & Juan {\bf tiene fr\'\i o}
\end{tabular}
\end{exquote}
We therefore argue that a one-to-one translation for such phrases is not
adequate but instead consider the highlighted phrases above as the correct
equivalences between the two languages. The task then, is to find a mechanism
for efficiently capturing regularities of this sort in the present framework.
There are a number of alternatives for achieving this. We will consider three.
\subsection{Lexical Neutralization}
\label{lex-neu-sec}
The first possibility for handling multi-lexeme regularities in LMT is to
eliminate support verbs from the SL transfer representation altogether, and to
reintroduce them during generation. In this case, a semantic representation for
the sentences must be proposed. For the sake of argument assume an
adjective-like intersective semantics for both the Spanish nouns {\em Juan} and
{\em sed} and the corresponding English noun and adjective:
\begin{quote}
\footnotesize
SL: john1$_{1}$ thirsty1$_{1}$\\
TL: juan1$_{1}$ sed1$_{1}$
\end{quote}
Then, the bilexicon would include, among other things:
\begin{exquote}
thirsty1$_{x}$ $\leftrightarrow$ sed1$_{x}$\\
hungry1$_{x}$ $\leftrightarrow$ hambre1$_{x}$\\
{\em etc.}
\end{exquote}
Lexicalist transfer would apply these equivalences to construct an appropriate
TL bag. During Spanish bag generation, the appropriate support verb (i.e. {\em
tener}) would be introduced by inspection of monolingual lexical information
associated with {\em sed} \acite{danlosetal92}, from which correct
instantiation of the orthography of the TL sentence would ensue. A variation of
this strategy would be to use a partially instantiated lexical sign
corresponding to the English support verb:
\begin{quote}
\footnotesize
\{ john1$_{1}$ , support-verb$_{2,1,3}$ , thirsty$_{3}$ \}
\end{quote}
During transfer, the support verb is translated as a partially instantiated
support verb in Spanish. The generation algorithm would then be applied such that
monolingual constraints in the Spanish grammar fully instantiated the semantics
and orthography of this verb according to the support verb requirements of its
complement noun.
\subsection{Lexical Variables}
\label{lex-var-sec}
The second mechanism for capturing multi-lexeme regularities assumes
translation variables similar to those used in several transfer systems
\acite{alshawietal92,bechetal91,russelletal91}. If one represents transfer
variables by {\tt tr({\em $<$restrictions$>$})}, then the necessary bilexical
entry would be:
\begin{exquote}
be1$_{x,y,z}$, {\tt tr(}Adj$_{z}${\tt )} $\leftrightarrow$
tener1$_{x,y,z}$, {\tt tr(}Noun$_{z}${\tt )}
\end{exquote}
This entry states that `be' translates as {\em tener} as long as its complement
adjective translates as the complement noun of {\em tener}. The transfer
algorithm is modified to accommodate the transfer variable by, for example,
recursively calling itself on the value of {\tt tr(}Adj$_{z}${\tt )}.
Generation, however, proceeds as before. A variation of this mechanism is to
use contextual rather than transfer variables. In this case, a particular
lexical context is specified which constraints translation equivalence in a
manner analogous to the way left and right contexts are used in morphological
rewriting rules \acite{kaplanetal94}. Thus, the transfer relation
\begin{exquote}
be1$_{x,y,z}$, (Adj$_{z}$) $\leftrightarrow$
tener1$_{x,y,z}$, (Noun$_{z}$)
\end{exquote}
would indicate that in the context of an adjective complement, `be' may
translate as {\em tener} or vice versa. The main difference between this and
the transfer variable variant is that the contextual elements, Adj and Noun,
can serve as context to multiple transfer relations within the same cover,
whereas this would not be possible with transfer variables. We will appeal to
contextual variables in Section \ref{bil-com-sec}.
The third mechanism uses bilingual lexical rules to map bilexical entries into
new bilexical entries. We have adopted this mechanism for certain multi-lexeme
translations because it allows the exploitation of monolingual lexical rules in
a motivated manner which integrates naturally with the LMT architecture, and
because it provides a framework in which to study differences between lexical
processes in different languages.
\section{Lexical and Bi-Lexical Rules}
The lexicon has taken a prominent place in several linguistic theories
\acite{pollardetal94,oehrleetal88}, not least because, given appropriate tools,
both general and idiosyncratic properties of language can be captured within a
uniform framework. Among the tools normally employed one finds lexical rules
\acite{dowty78,flickinger87,pollardetal94} and inheritance mechanisms
\acite{briscoeetal93a,flickingeretal92}. Lexical rules may be thought of as
establishing a relationship between lexical items such that given the presence
of one lexical item in the lexicon the existence of a further item may be
inferred. The regularities captured by lexical rules might include changes in
the subcategorization and control properties of a verb, the denotation of a
noun or the interpretation of a preposition. With the advent of lexically
oriented approaches to translation, it is worth considering whether and how the
generalizations captured by lexical rules might be exploited in MT.
In order to investigate this issue we have adopted the notion of a bi-lexical
rule. A bi-lexical rule \acite{trujillo92b,copestakeetal93} takes a bilexical
entry as input, and outputs a new bilexical entry. These rules may be seen as
expanding the bilexicon in order to increase its coverage; under this view,
they are somewhat analogous to lexical rules in that they reduce the number of
bilexical entries that need to be explicitly listed. Bi-lexical rules also
serve to capture lexical, syntactic and semantic regularities in the
translation between two languages by relating equivalent lexical processes
cross-linguistically.
\subsection{Simple Bi-lexical Rule}
We give a simple example of a bi-lexical rule before addressing the
multi-lexeme translations introduced earlier. Consider the relationship that
exists in English-Spanish translations between the translation of fruits and
the translation of their corresponding trees \acite{soleretal93}:
\begin{quote}
\footnotesize
\begin{center}
\begin{tabular}{|l|l| l |l|l|} \cline{1-2} \cline{4-5}
\multicolumn{2}{|c|}{\bf Fruit} & & \multicolumn{2}{c|}{\bf Tree}\\ \cline{1-2} \cline{4-5}
{\em English} & {\em Spanish} & & {\em English} & {\em Spanish}\\ \cline{1-2} \cline{4-5}
almond & almendra & & almond tree & almendro \\ \cline{1-2} \cline{4-5}
apple & manzana & & apple tree & manzano \\ \cline{1-2} \cline{4-5}
cherry & cereza & & cherry tree & cerezo \\ \cline{1-2} \cline{4-5}
orange & naranja & & orange tree & naranjo \\ \cline{1-2} \cline{4-5}
plum & ciruela & & plum tree & ciruelo \\ \cline{1-2} \cline{4-5}
lemon & lim\'on & & lemon tree & limonero \\ \cline{1-2} \cline{4-5}
\end{tabular}
\end{center}
\end{quote}
The relevant relationship may be described by the following bi-lexical
rule: \\
\begin{tabular}{cccc}
{\scriptsize
$\avmplus{\att{common-noun}\\
\attvaltyp{orth}{orth}\\
\attvalshrunktyp{syn}{syn}\\
\attvaltyp{qualia}{fruit1(x)}\\
\attvaltyp{lang}{english}\\
\attvaltyp{sem}{pred(x)}}$} & &
$\leftrightarrow$ &
{\scriptsize $\avmplus{\att{common-noun}\\
\attvaltyp{orth}{orth}\\
\attvalshrunktyp{syn}{syn}\\
\attvaltyp{qualia}{fruit1(x)}\\
\attvaltyp{lang}{spanish}\\
\attvaltyp{sem}{pred(x)}}$} \\
{\LARGE $\Downarrow$} {\scriptsize noun-noun} & & & {\LARGE $\Downarrow$} {\scriptsize fruit-tree}\\
{\scriptsize
$\avmplus{\att{common-noun}\\
\attvaltyp{orth}{orth}\\
\attvalshrunktyp{syn}{syn}\\
\attvaltyp{qualia}{fruit1(y)}\\
\attvaltyp{lang}{english}\\
\attvaltyp{sem}{pred(y,z)}}$}
&
{\scriptsize
$\avmplus{\att{common-noun}\\
\attvaltyp{orth}{tree}\\
\attvalshrunktyp{syn}{syn}\\
\attvaltyp{qualia}{tree1(z)}\\
\attvaltyp{lang}{english}\\
\attvaltyp{sem}{tree1(z)}}$}
&
$\leftrightarrow$ &
{\scriptsize $\avmplus{\att{common-noun}\\
\attvaltyp{orth}{orth + MORPH}\\
\attvalshrunktyp{syn}{syn}\\
\attvaltyp{qualia}{tree1(z)}\\
\attvaltyp{lang}{spanish}\\
\attvaltyp{sem}{pred(z)}}$}
\end{tabular}
\bigskip
This bi-lexical rule says that if there is a bilexical entry translating
English fruit nouns into Spanish fruit nouns, then there is a bilexical entry
translating `{\em noun} tree' in English into a morphologically derived
tree-denoting noun in Spanish.
We adopt Qualia structure \acite{pustejovsky91} as our lexical-semantic
representation formalism. According to Pustejovsky, Qualia structure is one of
the four main types of information to be associated with a lexical entry (the
others being Argument, Event and Inheritance structure). The information
incorporated in a Qualia structure specifies the semantics of a lexical item by
virtue of the relations and properties in which it participates. For this
example we assume a simplified Qualia value \acite{pustejovsky91} indicating
whether a noun denotes a tree or a fruit. Note that the morphology of the
output Spanish lexical sign is left implicit since it depends on the actual
noun used (see fruit-tree table above); in addition, the English rule mapping a
noun into a noun modifier is a practical simplification of the complex issue of
noun-noun modification which we do not address here
\acite{pustejovskyetal93,johnstonetal95}. Another
point to note is that we will be vague regarding the amount of information
shared between the input and output lexical signs of lexical rules; a full
treatment of this issue involves aspects of default unification which are
beyond the scope of this paper \acite{meurers94,lascaridesetal95}. Suffice it
to say that in our implementation, an attempt has been made to share maximum
information between input and output lexical signs, although values such
as semantic variables are not shared between input and output lexical signs.
In the abbreviated notation introduced earlier, the above bi-lexical rule will
be represented as:
\begin{quote}
\begin{tabular}{llcl}
\footnotesize
Ne$_{x}$ & & $\leftrightarrow$ & Ns$_{x}$ \\
$\Downarrow$ {\scriptsize identity} & & & $\Downarrow$ {\scriptsize fruit-tree}\\
Ne$_{y,z}$ & tree1$_{z}$ & $\leftrightarrow$ & Ns$'_{z}$
\end{tabular}
\end{quote}
Given the translation `apple -- {\em
manzana}', for example, the rule would operate as indicated below:
\begin{quote}
\footnotesize
\begin{tabular}{llcl}
apple1$_{x}$ & & $\leftrightarrow$ & manzana1$_{x}$ \\
$\Downarrow$ {\scriptsize identity} & & & $\Downarrow$ {\scriptsize fruit-tree}\\
apple1$_{y,z}$ & tree1$_{z}$ & $\leftrightarrow$ & manzano1$_{z}$
\end{tabular}
\end{quote}
Its output is the additional translation relation `apple tree - {\em manzano}'. Similar
translations are achieved for other fruits.
Clearly this rule should only apply to fruits which grow on trees and not to
fruits such as strawberries which are found on low growing plants. Such
restrictions need to be incorporated in the monolingual lexical signs and
rules.
Implementationally, bilexical rules may be applied off-line in order to expand
the bilexicon before processing, or they may be applied during transfer
to extend the bilexicon just sufficiently to enable transfer. We have
opted for the latter approach.
\subsection{Support Verbs}
We now show how bi-lexical rules can be used in the translation of `thirsty',
basing our analysis on the classification of support verbs proposed by
\cite{danlosetal92} for English-French translation. Their proposal, implemented
as part of a Eurotra project, involves transfer at the Interface Structure.
The essence of their approach is similar to that for multi-lexeme translations
given in Section \ref{lex-neu-sec}: the support verb is deleted from the SL
transfer structure, the adjective `thirsty' is translated into the TL noun
({\em sed} in our case), and an appropriate TL support verb is incorporated
into the TL sentence during generation. Information regarding which support
verb a noun requires is encoded in its lexical entry.
Support verbs can be of five types: neutral (e.g. `is thirsty'), durative (e.g.
`remain thirsty'), inchoative (e.g. `get thirsty'), terminative (e.g. `stop
being thirsty') and iterative (e.g. `be thirty again'). We will consider
neutral support verbs only although the other categories could also be handled
through bi-lexical rules. One difference between the present approach and that
of Danlos et~al.\ is that we equate the noun `thirst' with the noun {\em sed}
in the bilexicon, rather than equating an adjective and a noun, thus factoring
category and support verb differences:
\begin{exquote}
thirst1$_{x}$ $\leftrightarrow$ sed1$_{x}$
\end{exquote}
We believe this reflects more truly the translation relation that exists
between the two lexical items. An English-Spanish bi-lexical rule is then
introduced to derive the adjective on the English side and to include the
neutral support verb `be'; on the Spanish side the support verb {\em tener}, for
the noun {\em sed}, is introduced:
\begin{quote}
\footnotesize
\begin{tabular}{llcll}
& N$_{x}$ & $\leftrightarrow$ & & N[ntrl=tener]$_{x}$ \\
& $\Downarrow$ {\scriptsize adjective} & & & $\Downarrow$ {\scriptsize identity} \\
be1$_{e,s,y}$ & A[ntrl=be]$_{y}$ & $\leftrightarrow$ &
tener1$_{e,s,y}$ & N[ntrl=tener]$_{y}$
\end{tabular}
\end{quote}
Note that we underspecify the support verb for the input English noun to allow `John
has an unquenchable thirst' and similar examples. The neutral (ntrl) control
verb required by the English adjective is included in its lexical entry's
Qualia structure. Thus, a fuller TFS for `thirsty' is:
\begin{quote}
{\scriptsize
$\avmplus{\att{adjective}\\
\attvaltyp{orth}{thirsty}\\
\attvalshrunktyp{syn}{syn}\\
\attval{qualia}{\avmplus{\att{qualia}\\
\attval{supp-verbs}{\avmplus{\att{supp-verbs}\\
\attvaltyp{ntrl}{be(e,s,y)}\\
\attvaltyp{inch}{get(e,s,y)}}}}}\\
\attvaltyp{lang}{english}\\
\attvaltyp{sem}{thirsty1(y)}}$}
\end{quote}
In designing an appropriate Qualia structure we have added to the roles
proposed by \cite{pustejovsky91} (Constitutive, Formal, Telic and Agentive) in
order to incorporate information necessary for capturing particular phenomena
\acite{johnstonetal95}.
When translating `John is thirsty', the analyser constructs the
transfer representation:
\begin{quote}
\footnotesize
john1$_{1}$ be1$_{2,1,3}$ thirsty1$_{3}$
\end{quote}
We include the support verb `be' in our representation, even though it
has empty semantics, in order to encode scoping information -- i.e. to
prevent `John is a painter' translating as `a painter is John'; this rather
{\em ad hoc} solution could be replaced by a mechanism analogous to the labels
used in Underspecified Discourse Representation Theory \acite{reyle95,franketal95}.
During transfer, the bi-lexical rule above is applied to the bi-lexical entry
for `thirst' to yield:
\begin{quote}
be1$_{e,s,x}$, thirsty1$_{x}$ $\leftrightarrow$ tener1$_{e,s,x}$, sed1$_{x}$
\end{quote}
This multi-lexeme relation is used to translate `is thirsty' into {\em tiene
sed}; a separate entry translates `John' into {\em Juan}. Bag generation then
ensures that the TL bag yields a sentence which satisfies the constraints
specified by the TL grammar.
The intuitive description of the above process is that we consider `is thirsty'
not to be translatable compositionally, but instead to require a multi-lexeme
translation. The purpose of bi-lexical rules then is to minimize the
repetition of information in the bi-lexicon while allowing the exploitation of
monolingual lexical processes.
\subsection{Lexicalization Patterns}
\label{lex-pat-sec}
There are other translation phenomena which can be described through the use of
bi-lexical rules. Consider lexicalization patterns for example \acite{talmy85}:
\begin{exquote}
John {\bf swims across} the river.\\
Juan {\bf cruza} el r\'\i o {\bf nadando}.
\end{exquote}
In the English sentence, the main verb encodes manner (i.e. swimming) and motion,
while in Spanish it encodes path (i.e. across) and motion; the remaining
meaning component in each case is expressed through a modifier. Talmy
attributes these distinctions to differences in lexicalization patterns between
the two languages.
A previous approach to such translations has been to introduce the bilexical entries
`swim -- {\em nadar + ando}' and `across -- {\em cruzar}' \acite{beaven92b}.
This approach, however, only implicitly acknowledges that theses two
translations are only appropriate in conjunction, and that separately they
are in fact unintuitive. This not only increases the non-determinism of
transfer and generation, but can increase the likelihood of incorrect
translations for unseen sentences. In the bi-lexical rule view, one relates
verb translations to translations incorporating lexicalization patterns as follows:
\begin{quote}
\footnotesize
\begin{tabular}{llcll}
V$_{e,s}$ & & $\leftrightarrow$ & & V$'_{e,s}$ \\
$\Downarrow$ {\scriptsize identity} & & & & $\Downarrow$ {\scriptsize gerund} \\
V$_{f,t}$ & across1$_{f,x}$ & $\leftrightarrow$ & cruzar$_{f,t,x}$ & V[vform= ing]$'_{f,t}$
\end{tabular}
\end{quote}
This rule derives, for every (movement) verb translation, a multi-lexeme
translation which includes `across' as a modifier (we leave the restriction on
verbs to movement events implicit; also, a simplified description of `across'
is assumed \acite{trujillo95}).
Application of this rule to `swim -- {\em nadar}' may be depicted as follows:
\begin{quote}
\footnotesize
\begin{tabular}{llcll}
swim1$_{e,s}$ & & $\leftrightarrow$ & & nadar1$_{e,s}$ \\
$\Downarrow$ {\scriptsize identity} & & & & $\Downarrow$ {\scriptsize gerund} \\
swim1$_{f,t}$ & across1$_{f,x}$ & $\leftrightarrow$ & cruzar$_{f,t,x}$ & nadando1$_{f,t}$
\end{tabular}
\end{quote}
Lexicalist translation of `John swims across the river' can then proceed by
translating `swims across' with the output of this rule and the remaining elements
of the input via other bilexical entries.
\subsection{Head Switching}
The phenomenon of head switching in translation can be exemplified by the
following pair of sentences:
\begin{exquote}
John {\bf just} arrived.\\
Juan {\bf acaba de} llegar.
\end{exquote}
The problem with such translations is that the syntactic head in the SL
sentence is not the syntactic head in its translation. This is a major obstacle
for syntactic and even some semantic based translation systems because of the
recursive nature of their transfer representations.
Head switching has been given a number of solutions in a variety of systems
\acite{kaplanetal89,sadleretal91,russelletal91,whitelock92,kaplanetal93}.
In our framework, the solution is expressed by the following rule
\footnote{We ignore the (complex) issue of tense for this type of example of
head switching; we expect that it can be tackled independently of the present approach.}:
\begin{exquote}
\begin{tabular}{llcll}
& V$_{e,s}$ & $\leftrightarrow$ & & V$'_{e,s}$ \\
& $\Downarrow$ {\scriptsize identity} & & & $\Downarrow$ {\scriptsize infinitive} \\
just1$_{f}$ & V$_{f,t}$ & $\leftrightarrow$ & acabar\_de1$_{f,t,f}$ & V$'_{f,t}$
\end{tabular}
\end{exquote}
Application to the bilexical entry `arrive -- {\em llegar}' results in:
\begin{exquote}
just1$_{f}$ , arrive1$_{f,t}$ $\leftrightarrow$ acabar\_de1$_{f,t,f}$ , llegar1$_{f,t}$
\end{exquote}
Lexicalist translation progresses as before. To exemplify the use of bi-lexical
rules in head switching, we consider translation in embedded contexts in more
detail now. To translate between:
\begin{exquote}
Mary thinks John just arrived.\\
Mar\'\i a piensa que Juan acaba de llegar.
\end{exquote}
the parser constructs the following representation (again, ignoring issues of
scope and quantification):
\begin{exquote}
mary1$_{1}$ , think1$_{2,1,4}$ , john1$_{3}$ , just1$_{4}$ , arrive1$_{4,3}$
\end{exquote}
Assuming appropriate transfer of `Mary' and `John', translation of the embedded
clause obtains as follows. `Thinks' is translated by the following entry:
\begin{exquote}
think1$_{e,s,f}$ $\leftrightarrow$ pensar\_que1$_{e,s,f}$
\end{exquote}
In addition, the output of the previous bi-lexical rule serves for multi-lexeme transfer of
`just arrive' to give the incomplete bag:
\begin{exquote}
\{ pensar\_que1$_{2,1,4}$ , acabar\_de1$_{4,3,4}$ , llegar1$_{4,3}$ \}
\end{exquote}
The final result of transfer is the TL bag:
\begin{exquote}
\{ mar\'\i a1$_{1}$ , pensar\_que1$_{2,1,4}$ , juan1$_{3}$ , acabar\_de1$_{4,3,4}$ ,
llegar1$_{4,3}$ \}
\end{exquote}
During generation, {\em acabar de} is made the syntactic head of the sentence
through grammatical constraints in the Spanish grammar. Illustrative rules
might be:
\begin{exquote}
S$_{e,s}$ $\Rightarrow$ NP$_{s}$ VP$_{e,s}$\\
VP$_{e,s,c}$ $\Rightarrow$ Vvp$_{e,s,c}$ VP$_{c}$\\
VP$_{e,s,c}$ $\Rightarrow$ Vs$_{e,s,c}$ S$_{c}$
\end{exquote}
If {\em pensar\_que} has category Vs, and {\em acabar\_de} has category
Vvp, there is only one ordering of the TL bag by which the constraints
indicated by this small grammar can be satisfied, namely, the order given by
its translation:
\begin{exquote}
Mar\'\i a piensa\_que Juan acaba\_de llegar.
\end{exquote}
It may be noticed that head selection by the TL grammar is possible because the
event semantic constants in {\em acabar\_de} and {\em llegar} are the same. The
consequence of this is that modifiers which apply to `just arrived' and
`arrived' separately will be indistinguishable during TL generation. Avoiding
this problem entails transferring scoping domains for modifiers in order to
constraint generation. However, we have no readily implementable mechanism for
achieving this in LMT as yet.
This concludes our overview of the different translation mismatches that may be
handled through bi-lexical rules. We now consider some unresolved issues
arising from their use.
\section{Bi-lexical Rule Interaction}
\label{bil-com-sec}
One difficulty we have found with bilexical rules has been their composition.
For example, consider the following translation:
\begin{exquote}
1) John {\bf marched} the soldiers {\bf across} the valley.\\
1$'$) Juan le {\bf hizo cruzar} el valle a los soldados {\bf marchando}.
\end{exquote}
In our framework, two bi-lexical rules should be applied in such cases: one to
construct causative translations \acite{comrie85,levinetal95}:
\begin{exquote}
\begin{tabular}{llcll}
& march1 & $\leftrightarrow$ & & marchar1 \\
& $\Downarrow$ {\scriptsize causative} & & & $\Downarrow$ {\scriptsize infinitive} \\
& march1$_{causative}$ & $\leftrightarrow$ & hacer1 & marchar1$_{infinitive}$
\end{tabular}
\end{exquote}
The other to deal with differences in lexicalization patterns such as `march
across -- {\em cruzar marchando}'. The problem is that in isolation neither of
these rules could perform the above translation. Ideally one should be able to
use the output of one as input to the other to derive `march across -- {\em
hacer cruzar marchando}', but this is not possible because both bi-lexical
rules expect a mono-lexeme bilexical entry.
One possible solution is to manually add further bi-lexical rules which incorporate the
composition of other rules:
\begin{exquote}
\begin{tabular}{llcll}
V & & $\leftrightarrow$ & & V$'$ \\
$\Downarrow$ {\scriptsize causative} & & & & $\Downarrow$ {\scriptsize gerund} \\
V & across1 & $\leftrightarrow$ & hacer1 cruzar1 & V$'$
\end{tabular}
\end{exquote}
However, this solution leads to a combinatorial explosion in the number of
bi-lexical rules.
The line of work we are investigating combines bi-lexical rules with the
context variables given in Section \ref{lex-var-sec}. There remain problems in
our implementation, however, which will be evident from the following
description. In our proposed approach either the causative or the
lexicalization pattern bi-lexical rule, or both, incorporate a context variable
in their output bilexical entry. For example, assume that the variable is
included in the causative rule:
\begin{exquote}
\begin{tabular}{llcll}
V & & $\leftrightarrow$ & & V$'$ \\
$\Downarrow$ {\scriptsize causative} & & & & $\Downarrow$ {\scriptsize infinitive} \\
(V) & & $\leftrightarrow$ & hacer1 & (V$'$)
\end{tabular}
\end{exquote}
This rule says that whenever there is a verb bilexical entry, there is also
an entry which in the context of a causative verb introduces {\em hacer} in the
TL bag. Applying the rule to `march -- {\em marchar}' gives:
\begin{exquote}
\begin{tabular}{lllcll}
& march1 & $\leftrightarrow$ & & marchar1 \\
& $\Downarrow$ {\scriptsize causative} & & & $\Downarrow$ {\scriptsize infinitive} \\
2) & (march1$_{causative}$) & $\leftrightarrow$ & hacer1 & (marchar1$_{infinitive}$)
\end{tabular}
\end{exquote}
Lexicalist transfer of `march$_{causative}$ ... across ...' via the output of
this rule and that for lexicalization patterns proceeds as follow: the
causative reading of `march' unifies with the context lexical sign in 2) but
is not translated by it. The TL side therefore only contributes {\em hacer}
to the final TL bag. Via the bi-lexical rule given in Section
\ref{lex-pat-sec}, `march across' is transferred such that {\em cruzar} and {\em
marchando} form part of the final TL bag. The result is therefore {\em hacer
cruzar marchando}, which, in combination with the translation of the rest
of the sentence can form the basis for bag generation.
Our main problem is that of resolving conflicts between the syntactic
constraints imposed by each bi-lexical rule. The causative rule requires the
Spanish side to include an infinitive verb, while the lexicalization pattern
rule requires a gerundive verb. Clearly both constraints cannot be satisfied
for the same lexical sign {\em marchar1}. The problem reflects itself in our
proposal in that the rule which includes the contextual pattern must be chosen
carefully. If the lexicalization pattern rule rather than the causative rule
had included the contextual verb lexical sign, the gerundive {\em marchando}
could not have been generated. Instead, a sentence analogous to `John made the
soldiers march crossing the valley' would result, which is perhaps not
desirable. In other words, the conflict between gerundive and infinitive
morphology for `march' is decided manually in advance. The interaction of such
decisions with other bi-lexical rules therefore might be unpredictable, and
hence is left for further investigation.
\section{Implementation}
The implemented prototype system contains approximately 250 bilexical entries;
this figure includes 20 proper names, 20 multi-lexeme translations and 6
contextual rules. The following translations were done on a SUN Sparc
workstation using Allegro Common Lisp. The time taken to find all possible TL
sentences is given in seconds; total times are for CPU + typical garbage
collection times.
\begin{exquote}
\begin{tabular}{lr} \hline
Translation & Total (CPU) \\ \hline
John thinks Mary just arrived \\
Juan piensa\_que Mar\'\i a acaba\_de llegar & 50 (28) \\ \hline
John swam across the river\\
Juan cruz\'o el r\'\i o nadando & 19 (16) \\ \hline
John marched the soldiers \\
Juan hizo marchar a los soldados & 19 (17) \\ \hline
\end{tabular}
\end{exquote}
These timings are only intended to give some idea of the type and stage of our
implementation, rather than reflect the performance of an optimized system.
\section{Conclusion}
We have introduced the mechanism of bi-lexical rules for incorporating lexical
rules in MT. These rules establish correspondences between bilexical entries
such that given the presence of one entry, the existence of another bilexical
entry can be inferred. We presented various phenomena that can be described
using such rules: noun sense extensions, support verbs, lexicalization patterns
and head switching. The rules provide a useful and motivated extension to the
LMT paradigm by providing it with a uniform approach to the description of a number
of translation phenomena.
The problems arising from conflicting constraints imposed by different translation
relations are described, and a partial solution to these was offered involving
the combined use of bi-lexical rules and contextual variables.
Future work could consider implementing Mel'\v cuk's lexical functions
\acite{heylenetal94} in a manner similar to the way bi-lexical rules were used
in the translation of support verbs.
\section*{Acknowledgements}
Thanks to two anonymous reviewers for their valuable comments. The LKB was
implemented by Ann Copestake as part of the ESPRIT ACQUILEX project. Remaining
errors are mine.
\footnotesize
|
\section{INTRODUCTION}
The principal purpose of this paper is to provide a general treatment
of
two-body tau decays \cite{0}which only assumes Lorentz
invariance
and exploits the tree-like structure of the dominant contributions to
the $\tau ^{-}\tau ^{+}$
production-decay sequence. In particular, CP invariance and a
$(V\mp A)\ $%
structure of the tau charged-current is not assumed. In a separate
paper \cite{1}, it
has been reported that by means of the associated stage-two spin-
correlation
functions$\ $the scales of $\Lambda \approx \ few$\ $100GeV\
$can
be probed
at $M_Z$ center-of-mass energy in unpolarized $e^{-}e^{+}$\
collisions. The
scale of $1-2TeV\ $ can be probed at $10GeV$\ or $4GeV$.
Previously in the study of the weak-interaction's charged-current in
muonic
and in hadronic processes, it has been important to determine the
complete
Lorentz structure directly from experiment in a model independent
manner.
Here, in Sec. 2, eight semi-leptonic parameters are defined for a
specific
tau semi-leptonic decay mode such as $\tau ^{-}\rightarrow \rho ^{-
}\nu \ $.
The parameters are physically defined in terms of tau-decay
partial-width-intensities for polarized-final-states. They can also be
simply expressed in terms of the helicity amplitudes $A(\lambda
_\rho
,\lambda _\nu )\ $for $\tau ^{-}\rightarrow \rho ^{-}\nu \ $.
Besides model independence, a major current issue is whether or
not
there is an additional chiral coupling in the tau's charged-current. A
chiral classification of
additional structure is a natural phenomenological extension of the
symmetries of the standard $SU(2)_L\ X\ U(1)$\ electroweak lepton
model. The
requirement of $\bar u(p_\nu )\rightarrow \bar u(p_\nu )\frac
12(1+\gamma _5)
$ and/or $u(k_\tau )\rightarrow \frac 12(1-\gamma _5)u(k_\tau )$\
invariance
of the vector and axial current matrix elements $\langle \nu \left|
v^\mu
(0)\right| \tau \rangle $\ and $\langle \nu \left| a^\mu (0)\right| \tau
\rangle $,$\ $allows only $g_L,g_{S+P},g_{S^{-}+P^{-
},}g_{+}=f_M+f_E,$and $%
\tilde g_{+}=T^{+}+T_5^{+}\ $couplings. From this $SU(2)_L$
perspective, the
relevant experimental question is what are the best current limits on
such
additional couplings? Similarly, $\bar u(p_\nu )\rightarrow \bar
u(p_\nu
)\frac 12(1-\gamma _5)$ and/or $u(k_\tau )\rightarrow \frac
12(1+\gamma _5)\
u(k_\tau )$\ invariance selects the complimentary set of $%
g_R,g_{S-P},g_{S^{-}-P^{-},}g_{-}=f_M-f_E,$and $\tilde g_{-
}=T^{+}-T_5^{+}\ $%
couplings. The absence of $SU(2)_R$ couplings is simply built into
the
standard model; it is not predicted by it. So, what are the best
current
limits on such $SU(2)_R$ couplings in tau physics?
In Sec.3, as a step towards precision answers to these basic
questions, the
semi-leptonic parameters are expressed in terms of a ``$(V-A)+$\
additional
chiral coupling'' structure in the ${J^{Charged}}_{Lepton}$
current \cite{1}. Two tables display the resulting values of the
parameters when the
various additional chiral couplings $(g_i/2\Lambda _i)\ $are small
relative
to the standard $V-A$\ coupling $(g_L).$\
Sec. 4 gives the most general Lorentz invariant spin-correlation
functions
for $e^{-}e^{+}\rightarrow \tau ^{-}\tau ^{+}$\ followed by $\tau
\rightarrow \rho \nu ,a_1\nu ,K^{*}\nu $\ including both $\nu
_{L,R}$\
helicities and both $\bar \nu _{R,L}$\ helicities.\ These same
parameters
appear in the general angular distributions
\begin{equation}
\frac{dN}{d(\cos \theta_1^\tau )d(\cos \tilde \theta _a)d \tilde
\phi_a} = {\bf R}_{\pm \pm }
\end{equation}
for the polarized $\tau^{-}$ decay
chain, see Ref. \cite{C94}. So, they can also
be directly measured by means of longitudinally-polarized beams at
a
tau/charm factory or at a B-factory with longitudinally polarized
beams.
In Sec. 5, the two tests for leptonic CP violation in $\tau \rightarrow
\rho
\nu \ $ decay are generalized to $\tau \rightarrow a_1\nu \ $ decay
and to
two additional tests if there are $\nu _R\ $and $\bar \nu _L\
$couplings \cite{C94a}. Sec.6 treats $\tau ^{-}\rightarrow \pi ^{-
}\nu ,K^{-}\nu $\ decay. These modes each provide less
information
since here only two of the semi-leptonic parameters can be
measured. The fundamental $S^{-}$\ and $P^{-}$\ couplings do
not
contribute to $\tau \rightarrow \rho \nu ,a_1\nu ,K^{*}\nu $\ but
they are also found to be suppressed in $\tau ^{-}\rightarrow \pi ^{-
}\nu
,K^{-}\nu $\ decay$.$ In the appendix we list the $A(\lambda _\rho
,\lambda
_\nu )\ $for $\tau ^{-}\rightarrow \rho ^{-}\nu $\ for the most
general tau ${J^{Charged}}_{Lepton}$ current.
\section{PARAMETRIZATION OF TAU SEMI-LEPTONIC \protect\newline DECAY MODES}
The reader should be aware that it is not necessary to use the
helicity
formalism \cite{5} because the parameters introduced below will be
fundamentally defined in terms tau-decay partial width intensities
for
polarized-final-states. However, the helicity
formalism does provide a lucid, neat, and
flexible framework for connecting the most general Lorentz
invariant
couplings at the Lagrangian level, for describing tau lepton
decays, with the most general Lorentz invariant spin-correlation
functions for $\tau^- \tau^+$ pair production. In practice, the
helicity formalism also frequently provides insights and checks on
the
resulting formulas and their symmetries. We present the discussion
for the $\rho
\nu$ channel, but the
same formulas hold for the $a_1 \nu$ and $K^* \nu$ channels.
In the $\tau ^{-}$ rest frame, the
matrix element for $\tau ^{-}\rightarrow \rho ^{-}\nu$ is
\begin{equation}
\langle \theta _1^\tau ,\phi _1^\tau ,\lambda _\rho ,\lambda _\nu
|\frac
12,\lambda _1\rangle =D_{\lambda _1,\mu }^{\frac 12*}(\phi
_1^\tau ,\theta
_1^\tau ,0)A\left( \lambda _\rho ,\lambda _\nu \right)
\end{equation}
where $\mu =\lambda _\rho -\lambda _\nu $ and $\lambda_1$ is the
$\tau^{-}$
helicity. For the $CP$-conjugate
process, $\tau
^{+}\rightarrow \rho ^{+}\bar \nu \rightarrow \left( \pi ^{+}\pi
^o\right) \bar \nu $, in the $\tau ^{+}$ rest frame
\begin{equation}
\langle \theta _2^\tau ,\phi _2^\tau ,\lambda _{\bar \rho },\lambda
_{\bar
\nu }|\frac 12,\lambda _2\rangle =D_{\lambda _2,\bar \mu }^{\frac
12*}(\phi
_2^\tau ,\theta _2^\tau ,0)B\left( \lambda _{\bar \rho },\lambda
_{\bar \nu
}\right)
\end{equation}
with $\bar \mu =\lambda _{\bar \rho }-\lambda _{\bar \nu }$.
These formulas only assume Lorentz invariance and do not assume
any
discrete symmetry properties. Therefore, it is easy to use this
framework for
testing for the consequences of such addtional symmetries. In
particular, for $%
\tau ^{-}\rightarrow \rho ^{-}\nu $ and $\tau ^{+}\rightarrow \rho
^{+}\bar
\nu $ a specific discrete symmetry implies a specific relation among
the associated
helicity amplitudes:%
$$
\begin{array}{cc}
\underline{Invariance} & \underline{Relation} \\ P & A\left( -
\lambda _\rho
,-\lambda _\nu \right) =A\left( \lambda _\rho ,\lambda _\nu \right)
\\
& B\left( -\lambda _{\bar \rho },-\lambda _{\bar \nu }\right)
=B\left(
\lambda _{\bar \rho },\lambda _{\bar \nu }\right) \\
C & B\left( \lambda _{\bar \rho },\lambda _{\bar \nu }\right)
=A\left(
\lambda _{\bar \rho },\lambda _{\bar \nu }\right) \\
CP & B\left( \lambda _{\bar \rho },\lambda _{\bar \nu }\right)
=A\left(
-\lambda _{\bar \rho },-\lambda _{\bar \nu }\right) \\
\tilde T_{FS} & A^{*}\left( \lambda _\rho ,\lambda _\nu \right)
=A\left(
\lambda _\rho ,\lambda _\nu \right) \\
& B^{*}\left( \lambda _{\bar \rho },\lambda _{\bar \nu }\right)
=B\left(
\lambda _{\bar \rho },\lambda _{\bar \nu }\right) \\
CP\tilde T_{FS} & B^{*}\left( \lambda _{\bar \rho },\lambda
_{\bar
\nu
}\right) =A\left( -\lambda _{\bar \rho },-\lambda _{\bar \nu }\right)
\end{array}
$$
Measurement of a non-real helicity
amplitude implies
a violation of $\tilde T_{FS}$ invariance when a first-order
perturbation in an
``effective" hermitian Hamiltonian is reliable. So $\tilde T_{FS}$
invariance is
expected to be violated when there are significant final-state
interactions; and it is
to be distinguished from canonical $T$ invariance which requires
interchanging
``final'' and ``initial'' states, i.e. actual time-reversed reactions are
required.
\subsection{Definition by partial width intensities for
polarized-final-states}
The tau semi-leptonic decay parameters for $\tau ^{-}\rightarrow
\rho
^{-}\nu $, and likewise for $\tau ^{-}\rightarrow {a_1}^{-}\nu $
and
\newline $\tau
^{-}\rightarrow {K^*}\nu $ , are defined by
\begin{equation}
\begin{array}{c}
\zeta \equiv (\Gamma _L^{-}-\Gamma _T^{-})/(
{\cal S}_\rho \Gamma ) \\ \sigma \equiv (\Gamma _L^{+}-\Gamma
_T^{+})/(
{\cal S}_\rho \Gamma ) \\ \xi \equiv \frac 1\Gamma (\Gamma _L^{-
}+\Gamma
_T^{-})
\end{array}
\end{equation}
where $\Gamma \equiv \Gamma _L^{+}+\Gamma _T^{+}$ is the
total partial width
for $\tau ^{-}\rightarrow \rho ^{-}\nu $. The subscripts denote the
polarization of the final $\rho ^{-}$, either ``L=longitudinal'' or
``T=transverse'', and the superscripts denote ``$\pm $ for
sum/difference of
the $\nu _{L\ }$versus $\nu _R$ contributions''. Such final-state-
polarized
partial widths are in principle physical observables but their direct
measurement would require measurement of the polarizations of
both the final
$\rho ^{-}$ and $\nu $. In Sec. 4 below, we will explain how the
equivalent semileptonic parameters can be measured by various
spin-
correlation
techniques.
To be clear about the terminology and sign-conventions, note that in
terms of the helicity amplitudes $A(\lambda _\rho ,\lambda _\nu )$
these
final-state-polarized partial widths are:
\begin{equation}
\begin{array}{c}
\Gamma _L^{\pm }=\left| A(0,-\frac 12)\right| ^2\pm \left| A(0,\frac
12)\right| ^2 \\
\Gamma _T^{\pm }=\left| A(-1,-\frac 12)\right| ^2\pm \left| A(-
1,\frac
12)\right| ^2
\end{array}
\end{equation}
Recall \cite{C94} that by rotational invariance the other $\rho ^{-}$
helicity amplitudes are
forbidden; similarly for the $\rho ^{+}\
$mode in $%
\tau ^{+}\ $decay, the $B(1, \frac 12)$ and $B(-1,-\frac 12)$
amplitudes vanish.
To describe the contributions from the interference between the
longitudinal(%
$L$) and transverse($T$) vector-meson amplitudes in the decay
process, the
additional parameters are:
\begin{equation}
\begin{array}{c}
\omega \equiv I_{
{\cal R}}^{-}\ /({\cal R}_\rho \Gamma ) \\ \eta \equiv I_{
{\cal R}}^{+}\ /({\cal R}_\rho \Gamma ) \\ \omega ^{\prime
}\equiv
I_{
{\cal I}}^{-}\ /({\cal R}_\rho \Gamma ) \\ \eta ^{\prime }\equiv
I_{{\cal I}%
}^{+}\ /({\cal R}_\rho \Gamma )
\end{array}
\end{equation}
In terms of the helicity amplitudes these measurable $LT$-
interference
intensities are
\begin{equation}
\begin{array}{c}
I_{
{\cal R}}^{\pm }={\cal RE}\{A(0,-\frac 12)^{*}A(-1,-\frac 12)\pm
A(0,\frac
12)^{*}A(1,\frac 12)\} \\ =\left| A(0,-\frac 12)\right| \left| A(-1,-\frac
12)\right| \cos \beta _a \\
\pm \left| A(0,\frac 12)\right| \left| A(1,\frac 12)\right| \cos \beta
_a^R
\end{array}
\end{equation}
\begin{equation}
\begin{array}{c}
I_{
{\cal I}}^{\pm }={\cal IM}\{A(0,-\frac 12)^{*}A(-1,-\frac 12)\pm
A(0,\frac
12)^{*}A(1,\frac 12)\} \\ =\left| A(0,-\frac 12)\right| \left| A(-1,-\frac
12)\right| \sin \beta _a \\
\pm \left| A(0,\frac 12)\right| \left| A(1,\frac 12)\right| \sin \beta
_a^R
\end{array}
\end{equation}
where $\beta _a\equiv \phi _{-1}^a-\phi _0^a$, $\beta _a^R\equiv
\phi
_1^a-\phi _0^{aR}$\ are the measurable phase differences of of the
associated helicity amplitudes $A=\left| A\right| \exp \iota \phi $.\
Note that the hadronic factors ${\cal S}_\rho $ and ${\cal R}_\rho $
do
depend on the particular tau semi-leptonic decay channel. For the
$\rho $
mode they are given by
\begin{equation}
{\cal S}_\rho =\frac{1-2\frac{m_\rho ^2}{m^2}}{1+2\frac{m_\rho
^2}{m^2}}
\end{equation}
\begin{equation}
{\cal R}_\rho =\frac{\sqrt{2}\frac{m_\rho }m}{1+2\frac{m_\rho
^2}{m^2}}
\end{equation}
In this section, these ${\cal S}_\rho $ and ${\cal R}_\rho $ factors
have
been explicitly inserted into the definitions of the semi-leptonic
decay
parameters, so that quantities such as ${q_\rho }^2={m_\rho }^2$
can be
smeared over in application of the spin-correlation functions given
below. Such smearing is needed due to the finite $\rho $ width.
This
treatment assumes that the momentum
dependence (i.e. the dependence on ${q_\rho }^2$,$\ldots $ ) of the
form-factors $g_L$ and $g_i$ is negligible. Depending on the
application and
on the desired experimental test, more sophisticated treatments of
the
${%
q_\rho }^2$,$\ldots $ dependence should be used such as ones
which
incorporate results from recent QCD calculations for tau decays
\cite{6} and ones which include possible contributions
from addtional resonances such as the $\rho ^{\prime }$. Because
of
the
smearing and the good understanding of QCD in tau physics, we do
not expect
this to be a fundamental difficulty in practice but rather a technical
matter that
requires sufficient care.
For the $a_1,K^{*}\ $modes, replace respectively $m_\rho
\rightarrow
m_{a_1},m_{K^{*}}$. These factors numerically are $({\cal
S},{\cal R})_{\rho
,a_1,K^{*}}=0.454,0.445;-0.015,0.500;0.330,0.472.\ \ $Recall
\cite{7} that ${\cal S}_{\pi ,K}=1$\ for $J=0$, so ${\cal S}_{\rho
,a_1,K^{*}}$%
suppresses the spin signatures when $J\neq 0$. On the other hand,
${\cal R}%
_{\rho ,a_1,K^{*}}\ $doesn't appear for $J=0\ $channels since their
sequential-decay-chains end with the first stage.
Depending on the physics and/or experimental situation, it may
sometimes be
advantagous to rewrite the spin-correlation function(s) of interest
directly
in terms of the above final-state-polarized partial widths and
$LT$%
-interference intensities, instead of using the above $\tau $ semi-
leptonic
decay parameters.
For the CP conjugate modes, $\tau ^{+}\rightarrow \rho ^{+}\bar
\nu
$ and $%
\tau ^{+}\rightarrow {a_1}^{+}\bar \nu $, the formulas for their
semi-leptonic decay parameters are the same except that all
quantities are
``barred,'' and there is the substitution of helicity amplitudes $%
A(x,y)\rightarrow B(-x,-y)$. This parametrization only assumes
Lorentz
invariance; for example, a simple test of CP invariance is that each
''barred'' semi-leptonic parameter is measured to be equal to it's
``unbarred'' associate (within experimental
errors).
\section{Significance of semi-leptonic parameters versus \newline
``Chiral Couplings"}
The most general Lorentz coupling for \hskip
1em $\tau^{-
}\rightarrow \rho
^{-}\nu _{L,R}$ is
\begin{equation}
\rho _\mu ^{*}\bar u_{\nu _\tau }\left( p\right) \Gamma ^\mu
u_\tau
\left(
k\right)
\end{equation}
where $k_\tau =q_\rho +p_\nu $. It is convenient to treat the vector
and
axial vector matrix elements separately. In Eq.(11)
\begin{equation}
\Gamma _V^\mu =g_V\gamma ^\mu +
\frac{f_M}{2\Lambda }\iota \sigma ^{\mu \nu }(k-p)_\nu +
\frac{g_{S^{-}}}{2\Lambda }(k-p)^\mu +\frac{g_S}{2\Lambda
}(k+p)^\mu
+%
\frac{g_{T^{+}}}{2\Lambda }\iota \sigma ^{\mu \nu }(k+p)_\nu
\end{equation}
\begin{equation}
\Gamma _A^\mu =g_A\gamma ^\mu \gamma _5+
\frac{f_E}{2\Lambda }\iota \sigma ^{\mu \nu }(k-p)_\nu \gamma
_5
+
\frac{g_{P^{-}}}{2\Lambda }(k-p)^\mu \gamma
_5+\frac{g_P}{2\Lambda }%
(k+p)^\mu \gamma _5 +\frac{g_{T_5^{+}}}{2\Lambda }\iota
\sigma ^{\mu \nu
}(k+p)_\nu \gamma _5
\end{equation}
The parameter
$%
\Lambda =$ ``the scale of New Physics''. In effective field theory
this
is the scale at which new particle thresholds are expected to
occur or where the theory becomes non-perturbatively strongly-
interacting so as to overcome perturbative inconsistencies. In old-
fashioned renormalization theory $\Lambda$ is the scale at
which the calculational methods and/or the principles of
``renormalization''
breakdown, see for example \cite{th}. While some terms of the
above do occur as higher-order perturbative-corrections in the
standard model,
such SM
contributions are ``small'' versus the sensitivities of present tests in
$\tau$ physics
in the analogous cases of the $\tau$'s neutral-current and
electromagnetic-current
couplings, c.f.
\cite{d0}. For charged-current couplings, the situation should be
the
same.
Without additional theoretical, c.f. \cite{1}, or experimental
inputs, it is not possible to select what is the "best" minimal set of
couplings for
analyzing the structure of the tau's charged current. For instance,
by
Lorentz
invariance, there are the equivalence theorems that for the
vector
current%
\begin{eqnarray}
S\approx V+f_M, & T^{+}\approx -V+S^{-}
\end{eqnarray}
\noindent
and for the axial-vector current
\begin{eqnarray}
P\approx -A+f_E, & T_5^{+}\approx A+P^{-}
\end{eqnarray}
The matrix elements of the divergences of these charged-currents
are
\begin{equation}
(k-p)_\mu V^\mu =[g_V(m_\tau -m_\nu )
+
\frac{g_{S^{-}}}{2\Lambda }q^2+\frac{g_S}{2\Lambda }(m_\tau
^2-m_\nu ^2)
+%
\frac{g_{T^{+}}}{2\Lambda }(q^2-[m_\tau -m_\nu ]^2)]\bar u_\nu
u_\tau
\end{equation}
\begin{equation}
(k-p)_\mu A^\mu =[- g_A(m_\nu +m_\tau )
+
\frac{g_{P^{-}}}{2\Lambda }q^2+\frac{g_P}{2\Lambda }(m_\tau
^2-m_\nu ^2)
+%
\frac{g_{T_5^{+}}}{2\Lambda }(q^2-[m_\tau+m_\nu ]^2)]\bar
u_\nu \gamma
_5u_\tau
\end{equation}
Both the weak magnetism $\frac{f_M}{2\Lambda }$ and the weak
electricty $%
\frac{f_E}{2\Lambda }$ terms are divergenceless. On the other
hand, since $%
q^2=m_\rho ^2$, even when $m_\nu =m_\tau $ there are non-
vanishing
terms due to
the couplings $S^{-},T^{+},A,P^{-},T_5^{+}$.
\subsection{Semi-leptonic parameters' form in terms of $g_L$ plus
an \newline ``additional chiral coupling''}
We first display the expected forms for the above semi-leptonic
parameters
for the $\tau \rightarrow \rho \nu ,\ a_1\nu ,K^{*}\nu $ \ decay
modes for
the case of a pure $V-A$ chiral coupling as in the standard lepton
model. We assume that the mass of the tau neutrino and anti-
neutrino are negligible, see Table 1 in \cite{e4}. Next we will give
the form for the case of a single chiral coupling $%
(g_i/2\Lambda _i)$\ in addition to the standard $V-A$ coupling. In
this
case, we first list the formula for an arbitrarily large additional
contribution. In two separate tables we list the formulas assuming
that the
additional contribution is small versus the $V-A$ coupling.
Throughout this
paper, we usually suppress the entry in the ``$i$'' subscript on the
new-physics coupling-scale ``$\Lambda _i$'' when it is obvious
from
the
context of interest.
In the case of ``multi-additional'' chiral contributions, the general
formulas for $A(\lambda _\rho ,\lambda _\nu )$\ \ which are listed
in
the
appendix can be substituted into the above definitions so as to
derive
the
expression(s) for the ``multi-additional'' chiral contributions.
Frequently we will suppress the subscript on $m_{\tau}$.
{\em Pure }$V-A$ {\em coupling:}
\begin{equation}
\begin{array}{cc}
\zeta =\sigma =\omega =\eta =\xi =1 \\ \omega ^{\prime }= \eta
^{\prime }=0
\end{array}
\end{equation}
$V+A{\em \ also\ present:}$%
\begin{equation}
\begin{array}{cc}
\zeta = \xi & \omega =\xi \\ \sigma =1 & \eta =
1 \\ \xi =\frac{\left| g_L\right| ^2-\left| g_R\right| ^2}{%
\left| g_L\right| ^2+\left| g_R\right| ^2} & \omega ^{\prime }=\eta
^{\prime
}=0
\end{array}
\end{equation}
$S+P\ {\em also\ present:}$%
\begin{equation}
\zeta =\sigma =\left(
\begin{array}{c}
(1-2
\frac{m_\rho ^2}{m^2})\left| g_L\right| ^2+\frac m\Lambda [1-
\frac{m_\rho ^2%
}{m^2}]{\cal RE}(g_L^{*}g_{S+P}) \\ +\{\frac m{2\Lambda }[1-
\frac{m_\rho ^2}{%
m^2}]\}^2\left| g_{S+P}\right| ^2
\end{array}
\right) / ( {\cal S}_\rho {\cal D}^{+} )
\end{equation}
\begin{equation}
\xi =1
\end{equation}
\begin{equation}
\begin{array}{c}
\omega =\eta =
\sqrt{2}\frac{m_\rho }m\left( \left| g_L\right| ^2+\frac m{2\Lambda
}[1-%
\frac{m_\rho ^2}{m^2}]{\cal RE}(g_L^{*}g_{S+P})\right) / ( {\cal
R}_\rho {\cal D}^{+} ) \\ \omega
^{\prime }=\eta ^{\prime }=-\sqrt{2}\frac{m_\rho }{2\Lambda }[1-
\frac{m_\rho
^2}{m^2}]{\cal IM}(g_L^{*}g_{S+P})/ ( {\cal R}_\rho {\cal
D}^{+} )
\end{array}
\end{equation}
where
$$
{\cal D}^{+ }=(1+2\frac{m_\rho ^2}{m^2})\left| g_L\right|
^2+\frac
m\Lambda [1-%
\frac{m_\rho ^2}{m^2}]{\cal RE}(g_L^{*}g_{S+P})+\{\frac
m{2\Lambda }[1-\frac{%
m_\rho ^2}{m^2}]\}^2\left| g_{S+P}\right| ^2
$$
$S-P\ {\em also\ present:}$%
\begin{equation}
\zeta ,\sigma =\left( (1-2\frac{m_\rho ^2}{m^2})\left| g_L\right|
^2\mp
\{\frac m{2\Lambda }[1-(\frac{m_\rho ^2}{m^2})]\}^2\left| g_{S-
P}\right|
^2\right) / ( {\cal S}_\rho {\cal D}^{-} )
\end{equation}
where the upper(lower) sign on the ``rhs'' goes with the
first(second)
entry
on the ``lhs.''
\begin{equation}
\xi =1
\end{equation}
\begin{equation}
\omega =\eta =\sqrt{2}\frac{m_\rho }m\left| g_L\right| ^2/( {\cal
R}_\rho {\cal D}^{-} ),\ \
\omega ^{\prime }=\eta ^{\prime }=0
\end{equation}
where
$$
{\cal D}^{-}=(1+2\frac{m_\rho ^2}{m^2})\left| g_L\right|
^2+\{\frac
m{2\Lambda }[1-\frac{m_\rho ^2}{m^2}]\}^2\left| g_{S-P}\right|
^2
$$
$f_M+f_E\ {\em also\ present:}$
For this case we write the coupling constant of the sum of the weak
magnetism and the weak electricity couplings as
$$
g_{+}=f_M+f_E
$$
In this notation,
\begin{equation}
\zeta =\sigma =\left(
\begin{array}{c}
(1-2
\frac{m_\rho ^2}{m^2})\left| g_L\right| ^2+\frac{m_\rho
^2}{m\Lambda }{\cal %
RE}(g_L^{*}g_{+}) \\ +\frac{m_\rho ^2}{4\Lambda ^2}[-
2+\frac{m_\rho ^2}{m^2}%
]\left| g_{+}\right| ^2
\end{array}
\right) / ( {\cal S}_\rho {\cal D_T}^{+} )
\end{equation}
$$
\xi =1
$$
\begin{equation}
\begin{array}{c}
\omega =\eta =
\sqrt{2}\frac{m_\rho }m\left( \left| g_L\right| ^2-\frac m{2\Lambda
}[1+%
\frac{m_\rho ^2}{m^2}]{\cal RE}(g_L^{*}g_{+})+\frac{m_\rho
^2}{4\Lambda ^2}%
\left| g_{+}\right| ^2\right) / ( {\cal R}_\rho {\cal D_T}^{+} ) \\
\omega ^{\prime }=\eta
^{\prime }=-\frac{m_\rho }{\sqrt{2}\Lambda }[1-\frac{m_\rho
^2}{m^2}]{\cal IM%
}(g_L^{*}g_{+})/ ( {\cal R}_\rho {\cal D_T}^{+} )
\end{array}
\end{equation}
where
$$
{\cal D_T}^{+}=(1+2\frac{m_\rho ^2}{m^2})\left| g_L\right| ^2-
3\frac{m_\rho ^2%
}{m\Lambda }{\cal RE}(g_L^{*}g_{+})+\frac{m_\rho
^2}{4\Lambda ^2}[2+\frac{%
m_\rho ^2}{m^2}]\left| g_{+}\right| ^2
$$
$f_M-f_E\ {\em also\ present:}$
Similarly, we write the coupling constant of the difference of the
weak
magnetism and the weak electricity couplings as
$$
g_{-}=f_M-f_E
$$
and so,
\begin{equation}
\zeta ,\sigma =\left( (1-2\frac{m_\rho ^2}{m^2})\left| g_L\right|
^2\pm
\frac{m_\rho ^2}{4\Lambda ^2}\left| g_{-}\right| ^2\right) / ( {\cal
S}_\rho {\cal D}_T^{-} )
\end{equation}
where the upper(lower) sign on the ``rhs'' goes with the
first(second)
entry
on the ``lhs.''Also,
\begin{equation}
\xi =\left( (1+2\frac{m_\rho ^2}{m^2})\left| g_L\right| ^2-
3\frac{m_\rho ^2}{%
4\Lambda ^2}\left| g_{-}\right| ^2\right) / {\cal D}_T^{-}
\end{equation}
\begin{equation}
\omega ,\eta =\sqrt{2}\frac{m_\rho }m\left( \left| g_L\right| ^2\mp
\frac{%
m_\rho ^2}{4\Lambda ^2}\left| g_{-}\right| ^2\right) / ( {\cal
R}_\rho {\cal D}_T^{-} ) ,\ \
\omega ^{\prime }=\eta ^{\prime }=0\
\end{equation}
Here%
$$
{\cal D}_T^{-}=(1+2\frac{m_\rho ^2}{m^2})\left| g_L\right|
^2+3\frac{m_\rho
^2}{4\Lambda ^2}\left| g_{-}\right| ^2
$$
$T^{+}+T_5^{+}\ {\em also\ present:}$
We let
$$
\tilde g_{+}=g_{T+T_5}^{+}
$$
In this notation,
\begin{equation}
\zeta =\sigma =\xi=1
\end{equation}
Also
\begin{equation}
\omega =\eta =1 ;\ \ \omega ^{\prime }=\eta ^{\prime
}=0
\end{equation}
A single additional \ $\tilde g_{+}={g^{+}}_{T^{+}+T_5^{+}}\
$coupling does not
change the values from that of the pure $V-A$\ coupling.
$T^{+}-T_5^{+}\ {\em also\ present:}$
\begin{quotation}
We let
$$
\tilde g_{-}=g_{T-T_5}^{+}
$$
and so,
\begin{equation}
\zeta = \xi ,\ \ \sigma =1
\end{equation}
\begin{equation}
\xi =\frac{\left| g_L\right| ^2-\left| \frac{m\tilde g_{-}}{2\Lambda
}%
\right| ^2}{\left| g_L\right| ^2+\left| \frac{m\tilde g_{-}}{2\Lambda
}%
\right| ^2}
\end{equation}
\begin{equation}
\omega = \xi ,\ \ \eta =1,\ \omega ^{\prime }=\eta
^{\prime }=0\
\end{equation}
\end{quotation}
A single additional \ $\tilde g_{-}={g^{+}}_{T^{+}-T_5^{+}}\
$coupling is
equivalent to a single additional $V+A$\ coupling, except for the
interpretation of their respective chirality parameters.
\subsection{ Semi-leptonic parameters when ``additional chiral
coupling'' \newline is small}
In Table 1 for the $V+A$\ and for the $S\mp P$\ couplings, we list
the
``expanded forms'' of the above expressions for the case in which
there is a
single additional chiral coupling $(g_i/2\Lambda _i)$\ which is
small
relative to the standard $V-A$\ coupling $(g_L)$. Similarly, in
Table
2 is
listed the formulas for the additional tensorial couplings. The
tensorial
couplings include the sum and difference of the weak magnetism
and
electricity couplings, $g_{\pm }=f_M\pm f_E$, which involve the
momentum
difference $q_\rho =k_\tau -p_\nu $. The alternative tensorial
couplings $%
\tilde g_{\pm }={g^{+}}_{T^{+}\pm T_5^{+}}$ instead involve
$k_\tau +p_\nu $.%
Notice that except for the following coefficients the formulas
tablulated in
these two tables are short and simple. As above we usually suppress
the
entry in the ``$i$'' subscript on ``$\Lambda _i$.'' For \newline Table
1 these
coefficients are
\begin{equation}
\begin{array}{cc}
a=\frac{4m_\rho ^2}{m\Lambda }\frac{(1-\frac{m_\rho
^2}{m^2})}{(1-4\frac{%
m_\rho ^4}{m^4})} & d=\frac m{4\Lambda }(1-
\frac{m_\rho ^2}{m^2})\frac{(1-2\frac{m_\rho
^2}{m^2})}{(1+2\frac{m_\rho ^2}{%
m^2})} \\ b=\frac{m^2}{2\Lambda ^2}\frac{(1-\frac{m_\rho
^2}{m^2})^2}{(1-4%
\frac{m_\rho ^4}{m^4})} & e=
\frac{m^2}{4\Lambda ^2}\frac{(1-\frac{m_\rho
^2}{m^2})^2}{(1+2\frac{m_\rho ^2%
}{m^2})} \\ c=\frac{m_\rho ^2}{\Lambda ^2}\frac{(1-\frac{m_\rho
^2}{m^2})^2}{%
(1-4\frac{m_\rho ^4}{m^4})} & f=\frac m{2\Lambda }(1-
\frac{m_\rho ^2}{m^2})
\end{array}
\end{equation}
The coefficients for Table 2 are
\begin{equation}
\begin{array}{cc}
g=\frac{2m_\rho ^2}{m\Lambda }\frac{(1-4\frac{m_\rho
^2}{m^2})}{(1-4\frac{%
m_\rho ^4}{m^4})} & l=
\frac{m(1+9\frac{m_\rho ^2}{m^2}+2\frac{m_\rho
^4}{m^4})}{2\Lambda (1+2\frac{%
m_\rho ^2}{m^2})} \\ h=\frac{m_\rho ^2}{2\Lambda ^2}\frac{(1-
4\frac{m_\rho ^2%
}{m^2})}{(1-4\frac{m_\rho ^4}{m^4})} & m=
\frac{m_\rho ^2(2+\frac{m_\rho ^2}{m^2})}{2\Lambda
^2(1+2\frac{m_\rho ^2}{m^2%
})} \\ j=\frac{m_\rho ^2}{\Lambda ^2}\frac{(1-\frac{m_\rho
^2}{m^2})}{(1-4%
\frac{m_\rho ^4}{m^4})} & n=
\frac{m_\rho ^2(1-\frac{m_\rho ^2}{m^2})}{2\Lambda
^2(1+2\frac{m_\rho ^2}{m^2%
})} \\ k=\frac{3m_\rho ^2}{2\Lambda ^2(1+2\frac{m_\rho
^2}{m^2})} & o=\frac
m{2\Lambda }(1-\frac{m_\rho ^2}{m^2})
\end{array}
\end{equation}
Should experimental measurements indicate other than a pure
$g_L$
value of a semi-leptonic parameter, a smearing and more
sophisticated treated of these coefficients will be warrented.
Upon comparing the entries in these two tables, notice that (i) a
single
additional \ $\tilde g_{+}={g^{+}}_{T^{+}+T_5^{+}}\ $coupling
does not
change the values from that of the pure $V-A$\ coupling, and that
(ii) a
single additional \ $\tilde g_{-}={g^{+}}_{T^{+}-T_5^{+}}\
$coupling is
equivalent to a single additional $V+A$\ coupling, except for the
interpretation of their respective chirality parameters. This follows
as a consequence of Eqs.(14, 15) and the absence of contributions
from the $S^-$ and $P^-$ couplings to the $\rho$, $a_1$, and
$K^*$
modes.
We have displayed this equivalence in Table 2 to emphasize the fact
that the commonly assumed total absence of $\tilde g_{\pm}$
couplings in tau lepton decays is supported by tests of the
experimental/theoretical normalization of the decay rates, such as
by
universality tests in lepton physics; however, this assumption is not
directly supported by the empirical values of other semi-leptonic
decay parameters.
\section{SPIN-CORRELATION FUNCTIONS IN TERMS OF
THE
SEMI-LEPTONIC PARAMETERS}
\subsection{The full S2SC function}
For the production decay sequence $e^{-}e^{+}\rightarrow
Z^o,\gamma
^{*}\rightarrow \tau ^{-}\tau ^{+}\rightarrow (\rho ^{-}\nu )(\rho
^{+}\bar
\nu )$\ followed by $\rho ^{ch}\rightarrow \pi ^{ch}\pi ^o$\ the full
``Stage 2 Spin-Correlation'' function (S2SC) including both $\nu
_L$,\ $\nu
_R\ $helicities and both $\bar \nu _R$,\ $\bar \nu _L\ $helicities is
given by
\begin{eqnarray}
\begin{array}{c}
{\bf I}_7={\bf I(}E_1,E_2,\phi ;\tilde \theta _{a,}\tilde \phi _a;\tilde
\theta _{b,}\tilde \phi _b{\bf )} \\ =\stackunder{h_1,h_2}{\sum
}\left|
T(h_1,h_2)\right| ^2\ {\bf R}_{h_1,h_1}\ {\bf \bar R}_{h_2,h_2}\
\\ +e^{\iota
\phi }T(++)T^{*}(--){\bf r}_{+-}{\bf \bar r}_{+-}+e^{-\iota \phi
}T(--)T^{*}(++){\bf r}_{-+}{\bf \bar r}_{-+}
\end{array}
\end{eqnarray}
where $T(\lambda _1,\lambda _2)\ $are the production helicity
amplitudes
given in Ref. \cite{11} which describe $Z^o,\gamma
^{*}\rightarrow \tau ^{-}\tau ^{+}.$ This formula also holds if
either, or both, $\tau ^{\pm }\rightarrow
a_{_1}^{\pm }\nu \ $followed by $a_{_1}^{\pm }\rightarrow (3\pi
)^{\pm }.$
The specific $\tau ^{\mp }\ $decay channel determines which
``composite
decay density matrix'' ${\bf R}_{h_1,h_1}$,\ or $\ {\bf \bar
R}_{h_2,h_2}$\
, is to be inserted.
The literature on polarimetry methods and spin-correlation function
in tau physics includes Refs. \cite{a2, c1, C94, C94a, ch2}.
{\bf Formulas for }$\tau \rightarrow \rho \nu $:
Including both $\nu _L\ $and $\nu _R\ $helicities and using a
``compact
boldface formalism,'' we find \cite{C94a} the composite decay
density matrix for $\tau
^{-}\rightarrow \rho ^{-}\nu \rightarrow (\pi ^{-}\pi ^o)\nu $ is%
\begin{eqnarray}
{\bf R=}\left(
\begin{array}{cc}
{\bf R}_{++} & e^{\iota \phi _1^\tau }
{\bf r}_{+-} \\ e^{-\iota \phi _1^\tau }{\bf r}_{-+} & {\bf R}_{--}
\end{array}
\right)
\end{eqnarray}
In terms of the semi-leptonic parameters, the diagonal elements are
\begin{equation}
{\bf R}_{\pm \pm }={\bf n}_a[1\pm {\bf f}_a\cos \theta _1^\tau
]\mp (1/\sqrt{%
2})\sin \theta _1^\tau \sin 2\tilde \theta _a\ {\cal R}_\rho [\omega
\cos \tilde \phi
_a+\eta ^{\prime }\sin \tilde \phi _a]
\end{equation}
These give the angular distributions $ \frac{dN}{d(\cos \theta
_1^\tau )d(\cos \tilde \theta _a)d \tilde \phi_a} $ for the polarized
$\tau^{-}$ decay
chain, see Eq.(1) above. The off-diagonal elements depend on
\begin{eqnarray}
\begin{array}{c}
{\bf r}_{+-}=({\bf r}_{-+})^{*} \\ ={\bf n}_a{\bf f}_a\sin \theta
_1^\tau
+(1/\sqrt{2})\sin 2\tilde \theta _a\ {\cal R}_\rho {\cos \theta _1^\tau
[\omega \cos \tilde
\phi _a+\eta ^{\prime }\sin \tilde \phi _a]+\iota [\omega \sin \tilde
\phi
_a-\eta ^{\prime }\cos \tilde \phi _a]\ }
\end{array}
\end{eqnarray}
In Eqs.(40, 41),%
\begin{eqnarray}
\left(
\begin{array}{c}
{\bf n}_a \\ {\bf n}_a{\bf f}_a
\end{array}
\right) =\cos ^2\tilde \theta _a\frac{\Gamma _L^{\pm }}{\Gamma
}\pm
\frac 12\sin ^2\tilde \theta _a\frac{\Gamma _T^{\pm }}{\Gamma }
\end{eqnarray}
or equivalently%
\begin{eqnarray}
\begin{array}{c}
{\bf n}_a=\frac 1{8}(3+\cos 2\tilde \theta _a+\sigma {\cal S}_\rho
[1+3\cos 2\tilde
\theta _a]) \\ {\bf n}_a{\bf f}_a=\frac 1{8}(\xi [1+3\cos 2\tilde
\theta _a]+\zeta {\cal S}_\rho [3+\cos 2\tilde \theta _a])
\end{array}
\end{eqnarray}
Similarly, for the conjugate process $\tau ^{+}\rightarrow \rho
^{+}\bar
\nu \rightarrow (\pi ^{+}\pi ^o)\bar \nu \ $including both $\bar \nu
_R\ $and $%
\bar \nu _L\ $helicities
\begin{eqnarray}
{\bf \bar R=}\left(
\begin{array}{cc}
{\bf \bar R}_{++} & e^{\iota \phi _2^\tau }
{\bf \bar r}_{+-} \\ e^{-\iota \phi _2^\tau }{\bf \bar r}_{-+} & {\bf
\bar R}%
_{--}
\end{array}
\right)
\end{eqnarray}
In terms of the semi-leptonic parameters, the diagonal elements are
\begin{equation}
{\bf \bar R}_{\pm \pm }={\bf n}_b[1\mp {\bf f}_b\cos \theta
_2^\tau ]\pm (1/%
\sqrt{2})\sin \theta _2^\tau \sin 2\tilde \theta _b\ {\cal R}_\rho [\bar
\omega \cos \tilde
\phi _b-\bar \eta ^{\prime }\sin \tilde \phi _b]
\end{equation}
and%
\begin{eqnarray}
\begin{array}{c}
{\bf \bar r}_{+-}=({\bf \bar r}_{-+})^{*} \\ =-{\bf n}_b{\bf
f}_b\sin
\theta
_2^\tau -(1/\sqrt{2})\sin 2\tilde \theta _b\ {\cal R}_\rho {\cos \theta
_2^\tau [\bar
\omega \cos \tilde \phi _b-\bar \eta ^{\prime }\sin \tilde \phi
_b]+\iota
[\bar \omega \sin \tilde \phi _b+\bar \eta ^{\prime }\cos \tilde \phi
_b]\}
\end{array}
\end{eqnarray}
In Eqs.(45, 46),%
\begin{eqnarray}
\left(
\begin{array}{c}
{\bf n}_b \\ {\bf n}_b{\bf f}_b
\end{array}
\right) =\cos ^2\tilde \theta _b\frac{\bar \Gamma _L^{\pm }}{\bar
\Gamma }%
\pm \frac 12\sin ^2\tilde \theta _b\frac{\bar \Gamma _T^{\pm
}}{\bar
\Gamma }
\end{eqnarray}
or equivalently%
\begin{eqnarray}
\begin{array}{c}
{\bf n}_b=\frac 1{8}(3+\cos 2\tilde \theta _b+\bar \sigma {\cal
S}_\rho [1+3\cos
2\tilde \theta _b]) \\ {\bf n}_b{\bf f}_b=\frac 1{8}(\bar \xi [1+3\cos
2\tilde \theta _b]+\bar \zeta {\cal S}_\rho [3+\cos 2\tilde \theta _b])
\end{array}
\end{eqnarray}
{\bf Formulas for }$\tau \rightarrow a_1\nu $:
For the kinematic description of $\tau ^{-}\rightarrow a_1^{-}\nu
\rightarrow (\pi _1^{-}\pi _2^{-}\pi _3^{+})\nu $\ , the normal to
the
$(\pi
_1^{-}\pi _2^{-}\pi _3^{+})\ $decay triangle is used in place of the
$\pi
^{-}$\ momentum direction of the $\tau ^{-}\rightarrow \rho ^{-
}\nu
\rightarrow (\pi ^{-}\pi ^o)\nu \ $sequential decay \cite{bj}.
Including both $\nu _L\ $and $\nu _R\ $helicities, we find the
composite
decay density matrix for $\tau ^{-}\rightarrow a_1^{-}\nu
\rightarrow (\pi
_1^{-}\pi _2^{-}\pi _3^{+})\nu \ $ is%
\begin{equation}
{\bf R}^\nu =S_1^{+}{\bf R}^{+}+S_1^{-}{\bf R}^{-}
\end{equation}
where ${\bf R}^{\pm }$\ have the same the same form as the
earlier
matrix,
Eq.(39), except the elements now also have ``$\pm $'' superscripts,
see
below.\ $S_1^{\pm }\ $depend on the strong-interaction form-
factors
used to
describe the decay $a_1^{-}\rightarrow \pi _1^{-}\pi _2^{-}\pi
_3^{+}$.\
However, when the 3-body Dalitz plot is integrated over, only the
$S_1^{+}\ $%
term remains, so it can be absorbed into the overall normalization
factor
which removes any arbitrary form-factor dependence. In Eq.(49),
the
${\bf R}%
^{+}\ $composite decay matrix elements are%
\begin{eqnarray}
\begin{array}{c}
{\bf R}_{\pm \pm }^{+}=\{Eq.(40)\ with\ (1/\sqrt{2})\rightarrow (-
1/\sqrt{2}%
)\} \\ {\bf r}_{+-}^{+}=({\bf r}_{-+}^{+})^{*} \\ =\{Eq.(41)\ with\
(1/\sqrt{2}%
)\rightarrow (-1/\sqrt{2})\}
\end{array}
\end{eqnarray}
with%
\begin{eqnarray}
\left(
\begin{array}{c}
{\bf n}_a \\ {\bf n}_a{\bf f}_a
\end{array}
\right) =\sin ^2\tilde \theta _a\frac{\Gamma _L^{\pm }}{\Gamma
}\pm
(1-\frac 12\sin ^2\tilde \theta _a)\frac{\Gamma _T^{\pm
}}{\Gamma
}
\end{eqnarray}
or equivalently%
\begin{eqnarray}
\begin{array}{c}
{\bf n}_a=\frac 1{16}(10-2\cos 2\tilde \theta _a-\sigma {\cal
S}_\rho
[5+3\cos 2\tilde
\theta _a]) \\ {\bf n}_a{\bf f}_a=\frac 1{16}(-\xi [5+3\cos 2\tilde
\theta _a]+\zeta {\cal S}_\rho [10-2\cos 2\tilde \theta _a])
\end{array}
\end{eqnarray}
Similarly, the ${\bf R}^{-}\ $composite decay matrix elements are
\begin{equation}
{\bf R}_{\pm \pm }^{-}=-{\bf n}_a^{-}[1\mp {\bf f}_a^{-}\cos
\theta _1^\tau
]\mp (\sqrt{2})\sin \theta _1^\tau \sin \tilde \theta _a\ {\cal R}_\rho
[\eta \cos \tilde
\phi _a+\omega ^{\prime }\sin \tilde \phi _a]
\end{equation}
with%
\begin{eqnarray}
\left(
\begin{array}{c}
{\bf n}_a^{-} \\ {\bf n}_a^{-}{\bf f}_a^{-}
\end{array}
\right) =\cos \tilde \theta _a\frac{\Gamma _T^{\mp }}{\Gamma }
\end{eqnarray}
or%
\begin{eqnarray}
\begin{array}{c}
{\bf n}_a^{-}=\frac 12\cos \tilde \theta _a[\xi -\zeta {\cal S}_\rho ]
\\
{\bf n}_a^{-}%
{\bf f}_a^{-}=\frac 12\cos \tilde \theta _a[1-\sigma {\cal S}_\rho ]
\end{array}
\end{eqnarray}
Also%
\begin{eqnarray}
\begin{array}{c}
{\bf r}_{+-}^{-}=({\bf r}_{-+}^{-})^{*} \\ =\frac 12\sin \theta
_1^\tau \cos
\tilde \theta _a[1-\sigma {\cal S}_\rho ]+\sqrt{2}\sin \tilde \theta _a\
{\cal R}_\rho {\cos \theta
_1^\tau [\eta \cos \tilde \phi _a+\omega ^{\prime }\sin \tilde \phi
_a]+\iota [\eta \sin \tilde \phi _a-\omega ^{\prime }\cos \tilde \phi
_a]\}
\end{array}
\end{eqnarray}
For the conjugate process $\tau ^{+}\rightarrow a_1{}^{+}\bar \nu
\rightarrow (\pi _1^{+}\pi _2^{+}\pi _3^o)\bar \nu $\ ,%
\begin{equation}
{\bf \bar R}^{\bar \nu }=\bar S_1^{+}{\bf \bar R}^{+}+\bar
S_1^{-
}{\bf \bar R%
}^{-}
\end{equation}
The ${\bf \bar R}^{+}\ $matrix elements are%
\begin{eqnarray}
\begin{array}{c}
{\bf \bar R}_{\pm \pm }^{+}=\{Eq.(45)\ with\
(1/\sqrt{2})\rightarrow (-1/\sqrt{%
2})\} \\ {\bf \bar r}_{+-}^{+}=({\bf \bar r}_{-+}^{+})^{*} \\
=\{Eq.(46)\
with\ (1/\sqrt{2})\rightarrow (-1/\sqrt{2})\}
\end{array}
\end{eqnarray}
with%
\begin{eqnarray}
\left(
\begin{array}{c}
{\bf n}_b \\ {\bf n}_b{\bf f}_b
\end{array}
\right) =\sin ^2\tilde \theta _b\frac{\bar \Gamma _L^{\pm }}{\bar
\Gamma }%
\pm (1-\frac 12\sin ^2\tilde \theta _b)\frac{\bar \Gamma _T^{\pm
}}{\bar
\Gamma }
\end{eqnarray}
or%
\begin{eqnarray}
\begin{array}{c}
{\bf n}_b=\frac 1{16}(10-2\cos 2\tilde \theta _b-\bar \sigma {\cal
S}_\rho [5+3\cos
2\tilde \theta _b]) \\ {\bf n}_b{\bf f}_b=\frac 1{16}(-\bar \xi
[5+3\cos
2\tilde \theta _b]+\bar \zeta {\cal S}_\rho [10-2\cos 2\tilde \theta
_b])
\end{array}
\end{eqnarray}
The ${\bf \bar R}^{-}\ $matrix elements are
\begin{equation}
{\bf \bar R}_{\pm \pm }={\bf n}_b^{-}[1\pm {\bf f}_b^{-}\cos
\theta _2^\tau
]\mp \sqrt{2}\sin \theta _2^\tau \sin \tilde \theta _b\ {\cal R}_\rho
[\bar \eta \cos \tilde
\phi _b-\bar \omega ^{\prime }\sin \tilde \phi _b]
\end{equation}
and%
\begin{eqnarray}
\begin{array}{c}
{\bf \bar r}_{+-}^{-}=({\bf \bar r}_{-+}^{-})^{*} \\ =\frac 12\sin
\theta
_2^\tau \cos \tilde \theta _b[1-\bar \sigma {\cal S}_\rho
]+\sqrt{2}\sin \tilde \theta
_b\ {\cal R}_\rho {\cos \theta _2^\tau [\bar \eta \cos \tilde \phi _b-
\bar \omega ^{\prime
}\sin \tilde \phi _b]+\iota [\bar \eta \sin \tilde \phi _b+\bar \omega
^{\prime }\cos \tilde \phi _b]\}
\end{array}
\end{eqnarray}
with%
\begin{eqnarray}
\left(
\begin{array}{c}
{\bf n}_b^{-} \\ {\bf n}_b^{-}{\bf f}_b^{-}
\end{array}
\right) =\cos \tilde \theta _b\frac{\bar \Gamma _T^{\mp }}{\bar
\Gamma }
\end{eqnarray}
or%
\begin{eqnarray}
\begin{array}{c}
{\bf n}_b^{-}=\frac 12\cos \tilde \theta _b[\bar \xi -\bar \zeta {\cal
S}_\rho ] \\ {\bf n}%
_b^{-}{\bf f}_b^{-}=\frac 12\cos \tilde \theta _b[1-\bar \sigma {\cal
S}_\rho ]
\end{array}
\end{eqnarray}
\subsection{The simplest S2SC function}
The simpler 4 variable S2SC function including both $\nu \ $and
both $\bar
\nu $\ helicities is
\begin{eqnarray}
\begin{array}{c}
{\bf I}_4={\bf I(}E_1,E_2,\tilde \theta _1,\tilde \theta _2{\bf )} \\
=\left| T(+,-)\right| ^2 {\bf \rho }_{++} {\bf \bar \rho
}_{--}+\left|
T(-,+)\right| ^2 {\bf \rho }_{--} {\bf \bar \rho
}_{++}+\left| T(+,+)\right| ^2%
{\bf \rho }_{++} {\bf \bar \rho }_{++}+\left| T(-,-
)\right| ^2 {\bf \rho }_{--}%
{\bf \bar \rho }_{--}
\end{array}
\end{eqnarray}
This formula is in terms of the {\it integrated} composite decay
density
matrices for the $\tau ^{\pm }\rightarrow \rho ^{\pm }\nu \ $and/or
for the $%
\tau ^{\pm }\rightarrow a_1^{\pm }\nu $ decay chains with $\rho
^{\pm
}\rightarrow (2\pi )^{\pm }$\ and a$_1^{\pm }\rightarrow (3\pi
)^{\pm }$\ . Note that as for the $\bf R$'s in the preceding section,
here in Eq.(65) the $\rho$'s include both neutrino helicities. Here,
for convenience we suppress their ``boldface font''.
{\bf Formulas for }$\tau \rightarrow \rho \nu :$
For $\tau ^{-}\rightarrow \rho ^{-}\nu \rightarrow (\pi ^{-}\pi
^o)\nu
$,
with $\tau ^{-}$\ helicity $\lambda _1=h/2$%
\begin{eqnarray}
\begin{array}{c}
{\bf \rho }_{hh}\equiv \frac 1{\Gamma }\frac{dN}{d(\cos \theta
_1^\tau )d(\cos \tilde \theta _1)} \\ =\frac 18(3+\cos 2\tilde \theta
_1)S+\frac 1{16}(1+3\cos 2\tilde \theta _1)D
\end{array}
\end{eqnarray}
where
\begin{equation}
S=1+h\zeta {\cal S}_\rho \cos \theta _1^\tau
\end{equation}
\begin{equation}
D=-S(1-\cos 2\omega _1)+(\sigma {\cal S}_\rho +h\xi \cos \theta
_1^\tau )(1+3\cos 2\omega
_1)+h\omega {\cal R}_\rho 4\sqrt{2}\sin 2\omega _1\sin \theta
_1^\tau .
\end{equation}
Formulas for the Wigner rotation angles $\omega_{1,2}$ which are
solely functions respectively of $E_{1,2}$ are given in \cite{C94}.
It is important to note that if $\tilde \theta _1$\ is integrated out, i.e.
if the polarimetry information from the $\rho ^{-}\rightarrow (2\pi
)^{-}\ $%
stage is not included, then $D$ doesn't contribute. In this manner,
$\zeta $ is
measurable. Then inclusion of the $\tilde \theta _1$\ dependence
gives $D$ and
also enables separation of $\xi $ and $\omega $ because of their
differing
dependence on $\tilde \theta _1$%
{}.
For the CP conjugate process $\tau ^{+}\rightarrow \rho ^{+}\bar
\nu
\rightarrow (\pi ^{+}\pi ^o)\bar \nu $, with $\tau ^{+}$\ helicity
$\lambda
_1=h/2$
\begin{eqnarray}
\begin{array}{c}
{\bf \bar \rho }_{hh}\equiv \frac 1{\tilde \Gamma
}\frac{dN}{d(\cos
\theta
_2^\tau )d(\cos \tilde \theta _2)} \\ =\frac 18(3+\cos 2\tilde \theta
_2)\bar S+\frac 1{16}(1+3\cos 2\tilde \theta _2)\bar D
\end{array}
\end{eqnarray}
where
\begin{equation}
\bar S=1-h\bar \zeta {\cal S}_\rho \cos \theta _2^\tau
\end{equation}
\begin{equation}
\bar D=-\bar S(1-\cos 2\omega _2)+(\bar \sigma {\cal S}_\rho -h\bar
\xi \cos \theta
_2^\tau )(1+3\cos 2\omega _2)-h\bar \omega {\cal R}_\rho
4\sqrt{2}\sin 2\omega _2\sin
\theta _2^\tau .
\end{equation}
{\bf Formulas for }$\tau \rightarrow a_1\nu :$
For $\tau ^{-}\rightarrow a_1^{-}\nu \rightarrow (3\pi )^{-}\nu $,
with $%
\tau ^{-}$\ helicity $\lambda _1=h/2$
where%
\begin{eqnarray}
\begin{array}{c}
{\bf \rho }_{hh}\equiv \frac 1{\Gamma }\frac{dN}{d(\cos \theta
_1^\tau )d(\cos \tilde \theta _1)} \\ =\frac 14(3+\cos 2\tilde \theta
_1)S_{a_1}-\frac 1{32}(1+3\cos 2\tilde \theta _1)D_{a_1}
\end{array}
\end{eqnarray}
\begin{equation}
S_{a_1}=1+h\zeta {\cal S}_\rho \cos \theta _1^\tau
\end{equation}
\begin{equation}
D_{a_1}=S_{a_1}(3+\cos 2\omega _1)+(\sigma {\cal S}_\rho +h\xi
\cos \theta _1^\tau
)(1+3\cos 2\omega _1)+h\omega {\cal R}_\rho 4\sqrt{2}\sin
2\omega _1\sin \theta _1^\tau .
\end{equation}
The remarks above, following the analogous formulas in the $\rho $
case,
also apply here.
For the CP conjugate process $\tau ^{+}\rightarrow a_1^{+}\bar
\nu
\rightarrow (3\pi )^{+}\bar \nu $, with $\tau ^{+}$\ helicity
$\lambda _2=h/2
$
\begin{eqnarray}
\begin{array}{c}
{\bf \bar \rho }_{hh}\equiv \frac 1{\bar \Gamma }\frac{dN}{d(\cos
\theta
_2^\tau )d(\cos \tilde \theta _2)} \\ =\frac 14(3+\cos 2\tilde \theta
_2)\bar S_{a_1}-\frac 1{32}(1+3\cos 2\tilde \theta _2)\bar D_{a_1}
\end{array}
\end{eqnarray}
where
\begin{equation}
\bar S_{a_1}=1-h\bar \zeta {\cal S}_\rho \cos \theta _2^\tau
\end{equation}
\begin{equation}
\bar D_{a_1}=\bar S_{a_1}(3+\cos 2\omega _2)+(\bar \sigma {\cal
S}_\rho -h\bar \xi \cos
\theta _2^\tau )(1+3\cos 2\omega _2)-h\bar \omega {\cal R}_\rho
4\sqrt{2}\sin 2\omega
_2\sin \theta _2^\tau .
\end{equation}
\section{Tests for non-CKM-type leptonic CP violation}
By CP invariance each of the barred semi-leptonic parameters
should
equal, within experimental errors, its unbarred associate. However,
as was shown in Ref. \cite{C94}, if only $\nu_L$ and $\bar\nu_R$
exist, there are two simple tests for ``non-CKM-type" leptonic CP
violation in $\tau \rightarrow \rho \nu$ decay. Normally a CKM
leptonic-phase will contribute equally at tree level to both the
$\tau^-
$ and $\tau^+$ decay amplitudes (for exceptions see footnotes 14,
15 in Ref. \cite{C94}). These two tests follow because by CP
invariance $|B\left( \lambda_{\bar\rho},\lambda_{\bar\nu}\right) | =
|A\left(-\lambda_{\bar\rho},-\lambda_{\bar\nu}\right) |$. So the
two
tests for leptonic CP
violation are: %
\begin{equation}
\beta _a=\beta _b \hspace{2pc} {\bf first \hspace*{.4pc} test}
\end{equation}
where $\beta _a=\phi _{-1}^a-\phi _0^a$, $\beta _b=\phi _1^b-\phi
_0^b$, and
\begin{equation}
r_a=r_b \hspace{2pc} {\bf second \hspace*{.4pc} test}
\end{equation}
where%
\begin{equation}
r_a=\frac{|A\left( -1,-\frac 12\right) |}{|A\left( 0,-\frac 12\right)
|},r_b=%
\frac{|B\left( 1,\frac 12\right) |}{|B\left( 0,\frac 12\right) |}
\end{equation}
For sensitivity levels for $\tau \rightarrow \rho \nu$ decay, see Ref.
\cite{1}.
This analysis can be easily generalized \cite{C94a} to the $\tau
\rightarrow a_1\nu $ decay mode in which the
$a_1$ has
the opposite $CP$ quantum number to that of the $\rho $ : For the
$\tau ^{-}\rightarrow a_1^{-}\nu \rightarrow \left( \pi ^{-}\pi
^{-}\pi ^{+}\right) \nu ,\left( \pi ^o\pi ^o\pi ^{-}\right) \nu $ modes,
the composite-decay-density matrix is given by
\begin{equation}
\begin{array}{c}
\rho _{hh}=
\left( 1+h\cos \theta _1^\tau \right) \left[ \sin ^2\omega _1\cos
^2\tilde \theta _1 + ( 1- \frac 12\sin ^2\omega _1 ) \sin ^2\tilde
\theta
_1 \right] \\
+ \frac{r_a^2}2\left( 1-h\cos \theta _1^\tau \right) \left[ \left( 1+\cos
^2\omega_1\right) \cos ^2\tilde \theta _1 +\left( 1+\frac 12\sin
^2\omega
_1\right) \sin^2\tilde \theta _1 ] \\
-h\frac{r_a}{\sqrt{2}}\cos \beta _a\sin \theta
_1^\tau \sin 2\omega _1\left[ \cos ^2\tilde \theta _1-\frac
12\sin^2\tilde\theta_1\right]
\end{array}
\end{equation}
Table 3 shows that the sensitivity of the $a_1$ mode, versus that of
the $\rho$ mode, is about 2 times better for the $r_a$ measurement
and is about 5 times worse for the $\beta$ measurements. The
simpler $I_4$ function was used for $\sigma \left( r_a \right)$ and
the full $I_7$ was used for the other $\sigma$'s. The $CP$ and
$CP\tilde T_{FS}$ predictions for the phase relation between
$\beta
_a$ and $\beta _b$ are opposite, see Table 3 in \cite{e4}, so this
provides a method for distinguishing between a new physics effect
due to an unusual $CP$-violating final state interaction and one
with
a different mechanism of $CP$ violation.
It is also easy to generalize these simple tests so as to also include
$\nu
_R$ and $\bar
\nu
_L$ couplings. The necessary 4-variable S2SC is given by
\begin{equation}
\begin{array}{c}
I\left( E_\rho ,E_{\bar \rho },\tilde \theta _1,\tilde \theta _2\right)
\mid_{\nu _R,\bar \nu _L}=I_4 +\left( \lambda _R\right) ^2I_4\left(
\rho \rightarrow \rho ^R\right)
+\left( \bar \lambda _L\right) ^2I_4\left( \bar \rho \rightarrow
\bar\rho
^L\right) \\
+\left( \lambda _R\bar \lambda _L\right) ^2I_4\left( \rho \rightarrow
\rho^R,\bar \rho \rightarrow \bar \rho ^L\right)
\end{array}
\end{equation}
where $\lambda _R\equiv $ $\frac{|A\left( 0,\frac 12\right)
|}{|A\left(
0,-\frac 12\right) |},$ $\bar \lambda _L\equiv $ $\frac{|B\left( 0,-
\frac
12\right) |}{|B\left( 0,\frac 12\right) |}$ give the moduli's of the $\nu
_R$
and $\bar \nu _L$ amplitudes versus the standard amplitudes. The
corresponding
composite density matrices for $\tau \rightarrow \rho \nu $ with
$\nu _R$
and $\bar \nu _L$ final state particles are given by the substitution
rules:%
\begin{eqnarray}
\rho _{hh}^R=\rho _{-h,-h}\left( r_a\rightarrow r_a^R,\beta
_a\rightarrow
\beta _a^R\right) \\
\bar \rho _{hh}^L=\bar \rho _{-h,-h}\left( r_b\rightarrow
r_b^L,\beta
_b\rightarrow \beta _b^L\right)
\end{eqnarray}
where the $\nu _R$ and $\bar \nu _L$ moduli ratios and phase
differences
are defined by $r_a^R\equiv $ $\frac{|A\left( 1,\frac 12\right) |}{%
|A\left( 0,\frac 12\right) |},$ $r_b^L\equiv $ $\frac{|B\left( -1,-\frac
12\right) |}{|B\left( 0,-\frac 12\right) |},\beta _a^R\equiv \phi _1^a-
\phi
_0^{aR},\beta _b^L\equiv \phi _{-1}^b-\phi _0^{bL}$. The two
addtitional tests for ``non-CKM-type" leptonic CP violation if R-
handed $\nu$ and L-handed $\bar\nu$ exist are
\begin{equation}
{\beta _a}^R={\beta _b}^L \hspace{2pc} {\bf first \hspace*{.4pc}
\nu_R / {\bar\nu_L} \hspace*{.4pc} test}
\end{equation}
\begin{equation}
{r_a}^R={r_b}^L \hspace{2pc} {\bf second \hspace*{.4pc} \nu_R
/
{\bar\nu_L} \hspace*{.4pc} test}
\end{equation}
\section{DESCRIPTION OF $\tau ^{-}\rightarrow \pi ^{-}\nu
,K^{-
}\nu $}
The only observables for each of the $\tau ^{-}\rightarrow \pi ^{-
}\nu
,K^{-}\nu $\ modes which can be measured by spin-correlations are
the
chirality parameter $\xi _\pi =\frac{\left| A(-\frac 12)\right| ^2-\left|
A(\frac 12)\right| ^2}{\left| A(-\frac 12)\right| ^2+\left| A(\frac
12)\right| ^2}$ and the $\Gamma (\tau ^{-}\rightarrow \pi ^{-}\nu
)$,
\newline or $%
\Gamma (\tau ^{-}\rightarrow K^{-}\nu )$, partial width. The
relative phase
of the $A(\lambda _\nu )=A(\mp \frac 12)$\ amplitudes can not be
measured
unless, e.g. the $\nu _L$and $\nu _R$ have a common final decay
channel. For
$\tau ^{-}\rightarrow \pi ^{-}\nu ,$or $K^{-}\nu $,\ the $(k_\tau
+p_\nu )$\
effective couplings $(k_\tau +p_\nu )_\alpha \ V_{\nu \tau }^\alpha
$\ and $%
(k_\tau +p_\nu )_\alpha \ A_{\nu \tau }^\alpha $ are equivalent to
the
standard $q_{\pi ,\alpha }\ V_{\nu \tau }^\alpha $\ and $q_{\pi
,\alpha }\
A_{\nu \tau }^\alpha $\ couplings. Here $V_{\nu \tau }^\alpha $\
and $A_{\nu
\tau }^\alpha $\ are as in Eqs.(11-13). The $S^{-}$\ and $P^{-}$\
couplings can
contribute to the $\pi ^{-\text{\ }}$and $K^{-}$\ channels, whereas
they do
not for the $\rho ,a_{1,}K^{*}$\ modes. However, since $q\cdot
V\sim \frac{%
m_\pi ^2}{2\Lambda }g_{S^{-}}$\ and $q\cdot A\sim \frac{m_\pi
^2}{2\Lambda }%
g_{P^{-}}$\ their contribution is strongly suppressed for $\Lambda
>(\sim
1GeV)$\ scales.
By Lorentz invariance, there are the equivalence theorems that
$S^{-
}\approx
S\approx T^{+}\approx V$\ and P$^{-}\approx P\approx
T_5^{+}\approx A$\ .
The general helicity amplitudes for $\tau ^{-}\rightarrow \pi ^{-}\nu
,$or $%
K^{-}\nu $, for the above $q\cdot V$\ and $q\cdot A$\ couplings
are
\begin{equation}
\begin{array}{c}
A(\mp \frac 12)=g_L(E_\rho \pm q_\pi )
\sqrt{m_\tau (E_\nu \pm q_\pi )}+g_R(E_\rho \mp q_\pi
)\sqrt{m_\tau (E_\nu
\mp q_\pi )} \\ +(
\frac{m_\tau }{2\Lambda _i}%
)[g_{S+P}+g_{S-P}+(g_{S^{-}+P^{-}}+g_{S^{-}-P^{-
}})(\frac{m_\pi ^2}{m_\tau
^2-m_\nu ^2})]\{(E_\rho \pm q_\pi )\sqrt{m_\tau (E_\nu \pm q_\pi
)} \\
+(E_\rho \mp q_\pi )
\sqrt{m_\tau (E_\nu \mp q_\pi )}\} \\ +\tilde g_{+}(
\frac{m_\tau }{2\Lambda })\{(-1+\frac{m_\pi ^2}{m_\tau ^2-
m_\nu
^2})(E_\rho
\pm q_\pi )\sqrt{m_\tau (E_\nu \pm q_\pi )} \\ +(1+
\frac{m_\pi ^2}{m_\tau ^2-m_\nu ^2})(E_\rho \mp q_\pi
)\sqrt{m_\tau (E_\nu
\mp q_\pi )}\} \\ +\tilde g_{-}(
\frac{m_\tau }{2\Lambda })\{(1+\frac{m_\pi ^2}{m_\tau ^2-m_\nu
^2})(E_\rho
\pm q_\pi )\sqrt{m_\tau (E_\nu \pm q_\pi )} \\ +(-1+\frac{m_\pi
^2}{m_\tau
^2-m_\nu ^2})(E_\rho \mp q_\pi )\sqrt{m_\tau (E_\nu \mp q_\pi
)}\}
\end{array}
\end{equation}
The $\xi _\pi \ $ parameter can be measured by the stage-one
energy-correlation function $I(E_1^\pi ,E_2^\pi )$\ where $\rho
_{\pm \pm
}=1\pm \xi _\pi \cos \theta _1^\tau ,$ $\bar \rho _{\pm \pm }=1\mp
\bar \xi
_\pi \cos \theta _2^\tau $\ . From Eq(87) the effective $\lambda
=\left|
g_{eff}/g_L\right| ^2$\ value follows for
\begin{equation}
\frac{\Gamma (\tau \rightarrow \pi \nu _\tau )}{\Gamma (\pi
\rightarrow \mu
\nu _\mu )}=\frac \lambda 2\frac{m_\tau ^3}{m_\mu ^2m_\pi
}\left(
\frac{1-%
\frac{m_\pi ^2}{m_\tau ^2}}{1-\frac{m_\mu ^2}{m_\pi ^2}}\right)
^2
\end{equation}
For example,
$$
\begin{array}{cc}
\lambda _{S+P}=\left| 1+\frac{m_\tau }{2\Lambda
}\frac{g_{S+P}}{g_L}\right|
^2, & \lambda _{\tilde g_{+}}=\left| 1-\frac{m_\tau }{2\Lambda
}\frac{\tilde
g_{+}}{g_L}(1-\frac{m_\pi ^2}{m_\tau ^2})\right| ^2
\end{array}
$$
\begin{center}
{\bf Acknowledgments}
\end{center}
For helpful discussions, we thank experimentalists and theorists at
Cornell,
DESY, Valencia, and at the Montreux workshop. This work was
partially
supported by U.S. Dept. of Energy Contract No. DE-FG 02-
96ER40291.
\section*{Appendix: The helicity amplitudes in terms of the chiral
couplings}
In Sec. 2, the simple symmetry relations among the amplitudes are
possible
because of the Jacob-Wick phase conventions that were built into
the helicity formalism \cite{5}. In combining these amplitudes with
results from calculations of
similar amplitudes by diagramatic methods, care must be exercised
to insure that the same phase
conventions are being used (c.f. appendix in \cite{11}).
The helicity amplitudes for $\tau^{-
}\rightarrow \rho
^{-}\nu _{L,R}$ for both $(V\mp A)$ couplings and
$%
m_\nu $ arbitrary are for $\nu _L$ so $\lambda _\nu =-\frac 12$,%
\begin{eqnarray}
A\left( 0,-\frac 12\right) & = & g_L
\frac{E_\rho +q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu
+q_\rho
\right) } -g_R
\frac{E_\rho -q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu -
q_\rho
\right) } \\ A\left( -1,-\frac 12\right) & = & g_L
\sqrt{2m_\tau \left( E_\nu +q_\rho \right) } -g_R\sqrt{2m_\tau \left(
E_\nu -q_\rho \right) }.
\end{eqnarray}
and for $\nu _R$ so $\lambda _\nu =\frac 12$,%
\begin{eqnarray}
A\left( 0,\frac 12\right) & = & -g_L
\frac{E_\rho -q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu -
q_\rho
\right) } +g_R
\frac{E_\rho +q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu
+q_\rho
\right) } \\ A\left( 1,\frac 12\right) & = & -g_L
\sqrt{2m_\tau \left( E_\nu -q_\rho \right) } +g_R\sqrt{2m_\tau \left(
E_\nu +q_\rho \right) }
\end{eqnarray}
Note that
$g_L,g_R$
denote the
`chirality' of the coupling and $\lambda _\nu =\mp \frac 12$ denote
the
handedness of $\nu _{L,R}$. For $(S \pm P)$ couplings, the
additional contributions are
\begin{eqnarray}
A(0,-\frac 12) & =g_{S+P}(
\frac{m_\tau }{2\Lambda })\frac{2q_\rho }{m_\rho }\sqrt{m_\tau
(E_\nu
+q_\rho )} +g_{S-P}(\frac{m_\tau }{2\Lambda })\frac{2q_\rho
}{m_\rho }%
\sqrt{m_\tau (E_\nu -q_\rho )}, \quad
A(-1,-\frac 12) & =0
\end{eqnarray}
\begin{eqnarray}
A(0,\frac 12) & =g_{S+P}(
\frac{m_\tau }{2\Lambda })\frac{2q_\rho }{m_\rho }\sqrt{m_\tau
(E_\nu
-q_\rho )} +g_{S-P}(\frac{m_\tau }{2\Lambda })\frac{2q_\rho
}{m_\rho }%
\sqrt{m_\tau (E_\nu +q_\rho )}, \quad
A(1,\frac 12) & =0
\end{eqnarray}
The two types of tensorial couplings, $g_\pm = f_M \pm f_E$ and
$\tilde{g}_{\pm}={g^+}_{T^+ \pm T_5^+}$, give the additional
contributions
\begin{eqnarray*}
A\left( 0,\mp\frac 12\right) & = & \mp g_{+} (
\frac{m_\tau }{2\Lambda }) \left[
\frac{E_\rho \mp q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu
\pm q_\rho
\right) } - \frac{m_\nu }{m_\tau }
\frac{E_\rho \mp q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu \mp
q_\rho
\right) } \right] \\
& & \pm g_{-} (
\frac{m_\tau }{2\Lambda }) \left[
- \frac{m_\nu }{m_\tau }
\frac{E_\rho \pm q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu \pm
q_\rho
\right) } + \frac{E_\rho \pm q_\rho }{m_\rho } \sqrt{m_\tau \left(
E_\nu
\mp q_\rho
\right) } \right] \\
& & \mp \tilde g_{+} (
\frac{m_\tau }{2\Lambda }) \left[
\frac{E_\rho \pm q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu
\pm q_\rho
\right) } + \frac{m_\nu }{m_\tau }
\frac{E_\rho \mp q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu \mp
q_\rho
\right) } \right] \\
& & \pm \tilde g_{-} (
\frac{m_\tau }{2\Lambda }) \left[
\frac{m_\nu }{m_\tau }
\frac{E_\rho \pm q_\rho }{m_\rho } \sqrt{m_\tau \left( E_\nu \pm
q_\rho
\right) } + \frac{E_\rho \mp q_\rho }{m_\rho } \sqrt{m_\tau \left(
E_\nu
\mp q_\rho
\right) } \right]
\end{eqnarray*}
\begin{eqnarray*}
A\left( \mp 1,\mp\frac 12\right) & = & \mp \sqrt{2} g_{+} (
\frac{m_\tau }{2\Lambda }) \left[
\sqrt{m_\tau \left( E_\nu
\pm q_\rho
\right) } - \frac{m_\nu }{m_\tau }
\sqrt{m_\tau \left( E_\nu \mp
q_\rho
\right) } \right] \\
& & \pm \sqrt{2} g_{-} (
\frac{m_\tau }{2\Lambda }) \left[
- \frac{m_\nu }{m_\tau }
\sqrt{m_\tau \left( E_\nu \pm
q_\rho
\right) } + \sqrt{m_\tau \left( E_\nu
\mp q_\rho
\right) } \right] \\
& & \mp \sqrt{2} \tilde g_{+} (
\frac{m_\tau }{2\Lambda }) \left[
\sqrt{m_\tau \left( E_\nu
\pm q_\rho
\right) } + \frac{m_\nu }{m_\tau }
\sqrt{m_\tau \left( E_\nu \mp
q_\rho
\right) } \right] \\
& & \pm \sqrt{2} \tilde g_{-} (
\frac{m_\tau }{2\Lambda }) \left[
\frac{m_\nu }{m_\tau }
\sqrt{m_\tau \left( E_\nu \pm
q_\rho
\right) } + \sqrt{m_\tau \left( E_\nu
\mp q_\rho
\right) } \right]
\end{eqnarrray*}
|
\section{#1}}
\newcommand{\:\mbox{\sf Z} \hspace{-0.82em} \mbox{\sf Z}\,}{\:\mbox{\sf Z} \hspace{-0.82em} \mbox{\sf Z}\,}
\newcommand{\mbox{\scriptsize \sf Z} \! \! \mbox{\scriptsize \sf Z}}{\mbox{\scriptsize \sf Z} \! \! \mbox{\scriptsize \sf Z}}
\def\Multb#1#2{\left[#1 \atop #2 \right]}
\def\Mult#1#2{\biggl[{#1 \atop #2}\biggr]}
\def\Mults#1#2{\left[{\textstyle {#1 \atop #2} } \right]}
\def\mbox{e}{\mbox{e}}
\def\mbox{\scriptsize e}{\mbox{\scriptsize e}}
\def\case#1#2{{\textstyle{#1\over #2}}}
\def\varepsilon{\varepsilon}
\def\lambda{\lambda}
\def\mod#1{\; (\bmod \: #1)}
\begin{document}
\title{Fermionic solution of the Andrews-Baxter-Forrester model II:
proof of Melzer's polynomial identities}
\author{S.~Ole Warnaar\thanks{
e-mail: {\tt <EMAIL>}}
\\
Mathematics Department\\
University of Melbourne\\
Parkville, Victoria 3052\\
Australia}
\date{August, 1995 \\ \hspace{1mm}
\\
Preprint No. 17-95}
\maketitle
\begin{abstract}
We compute the one-dimensional configuration sums
of the ABF model using the fermionic technique
introduced in part I of this paper.
Combined with the results of Andrews,
Baxter and Forrester, we find proof of polynomial
identities for finitizations of the
Virasoro characters $\chi_{b,a}^{(r-1,r)}(q)$ as
conjectured by Melzer.
In the thermodynamic limit these identities
reproduce Rogers--Ramanujan type identities
for the unitary minimal Virasoro characters,
conjectured by the Stony Brook group.
We also present a list of additional Virasoro character identities
which follow from our proof of Melzer's identities and application
of Bailey's lemma.
{\bf Key words:} ABF model, One-dimensional configuration sums;
Fermi lattice-gas; Melzer's polynomial identities;
Rogers--Ramanujan identities; Virasoro characters.
\end{abstract}
\newpage
\nsection{Introduction}
Probably among the most celebrated results in
mathematics are the identities of Rogers and
Ramanujan~\cite{Rogers94,Rogers19,Ramanujan}
\begin{equation}
\sum_{m = 0}^{\infty} \frac{q^{m(m+a)}}{(q)_{m}}
=\frac{1}{(q)_{\infty}}\sum_{j=-\infty}^{\infty} (-1)^j
q^{j(5j +1+2a)/2} \qquad \qquad a=0,1,
\label{RR}
\end{equation}
where
$(q)_m = \prod_{k=1}^m (1-q^k)$, $m>0$ and
$(q)_0=1$.
In the context of modern physics, one recognizes the right-hand side
of these identities to be the Rocha-Caridi expression for the Virasoro
characters $\chi^{(2,5)}_{1,2-a}(q)$ of minimal conformal field theory
$M(2,5)$~\cite{Rocha}.
As such, the Rogers--Ramanujan identities can be seen
as character identities of some Virasoro algebra. A natural question
is whether the other Virasoro characters also admit
identities of the Rogers--Ramanujan type.
For the important
class of {\em unitary} minimal models $M(r-1,r)$,
this was answered affirmative in a remarkable paper by the
Stony Brook group~\cite{KKMMa}.\footnote{By now character identities
of Rogers--Ramanujan type for all minimal Virasoro characters
$\chi^{(p,p')}(q)$ have been found~\cite{KKMMa,FQ,BM}.}
However, the results of ref.~\cite{KKMMa}
were all based on extensive numerical studies, and actual
proofs remained elusive.
Among the many methods of proof of the original Rogers--Ramanujan
identities an elegant approach is that of first proving
the polynomial identities \cite{Schur,Andrews}
\begin{equation}
\sum_{m=0}^{\infty} q^{m(m+a)}
\Multb{L-m-a}{m} =
\sum_{j=-\infty}^{\infty} (-1)^j
q^{j(5j+1+2a)/2}
\Multb{L}{\lfloor \frac{1}{2}(L -5j-a)\rfloor},
\label{finiteRR}
\end{equation}
for all $L\geq a$.
Here $\lfloor x \rfloor$ denotes the integer part of $x$
and
$\Mults{N}{m}$ is a Gaussian polynomial
defined as
\begin{equation}
\renewcommand{\arraystretch}{1.5}
\Mult{N}{m} = \left\{
\begin{array}{ll}
\displaystyle \frac{(q)_N}{(q)_{m}(q)_{N-m}} \qquad & 0\leq m \leq N \\
0& \mbox{otherwise.}
\end{array}
\right.
\label{qpoly}
\end{equation}
Clearly, in the limit of $L\to\infty$ we recover the Rogers--Ramanujan
identity (\ref{RR}).
To proof the {\em finitized} Rogers--Ramanujan identities (\ref{finiteRR})
it suffices to check that both left- and right-hand side satisfy
the elementary recurrences $\mbox{$f_L=f_{L-1}+q^{L-1} f_{L-2}$}$
as well as the same initial conditions for $L=a,a+1$.
In an attempt to find proofs of the
identities for the characters $\chi^{(r-1,r)}_{b,a}(q)$ (see next
section for their actual form), Melzer followed Schur's approach
and conjectured
finitizations similar to those in (\ref{finiteRR}).
However, Melzer's polynomial identities were sufficiently complicated
not to lead to a straightforward proof using recurrences.
It was only after Melzer proved the cases
$r=3$ (Ising) and $r=4$ (tricritical Ising)~\cite{Melzer}
that Berkovich succeeded in proving recurrences for the polynomial
identities for all $\chi^{(r,r-1)}_{b,1}(q)$~\cite{Berkovich}.
In this paper we present a combinatorial proof for
Melzer's identities, based on yet another observation made by
Melzer. Again the motivation for this has been the original
Rogers-Ramanujan identities (\ref{RR}), whose finitization
(\ref{finiteRR}) can be viewed as evaluations of the sum
\begin{equation}
\renewcommand{\arraystretch}{1}
\sum_{
\begin{array}{c} \scriptstyle
\sigma_1,\ldots,\sigma_{L-1}=0,1\\
\scriptstyle
\sigma_j\sigma_{j+1}=0
\end{array}}
q^{ \;\; \displaystyle \sum_{k=1}^{L-1} k\sigma_k} \qquad
\qquad \sigma_0=a, \; \sigma_L=0,
\label{HHM}
\end{equation}
in two intrinsically different ways.
Similar to this, Melzer has argued that the polynomial identities for
the finitized $\chi^{(r-1,r)}_{b,a}(q)$ characters arise from
computing the sums
\begin{equation}
\renewcommand{\arraystretch}{1}
X_L(a,b) = \!\!\!\sum_{
\begin{array}{c} \scriptstyle
\sigma_1,\ldots,\sigma_{L-1}=0 \\
\scriptstyle
|\sigma_{j+1}\!-\!\sigma_j|=1
\end{array}}^{r-2} \!\!
q^{\;\;\displaystyle \sum_{k=1}^L
k|\sigma_{k+1}-\sigma_{k-1}|/4}
\quad \sigma_0=a-1, \; \sigma_L=b-1, \; \sigma_{L+1}=b,
\label{confsums}
\end{equation}
for all $a=1,\ldots,r-1$ and $b=1,\ldots,r-2$.
We will take this observation as the starting point for proving
the polynomial and Rogers--Ramanujan identities for the
(finitized) characters $\chi^{(r-1,r)}_{b,a}(q)$.
That is, we give two different methods to compute (\ref{confsums}),
one leading to a so called {\em fermionic} expression similar to
the left-hand side of (\ref{finiteRR}) and one method leading to a
so-called {\em bosonic} expression similar to the right-hand side
of (\ref{finiteRR}).
In fact, it should be noted that $X_L(a,b)$ defined above is
exactly the {\em one-dimensional configuration sum} $X_L(a,b,c)$, with
$c=b+1$, as defined by Andrews, Baxter and Forrester in their computation
of the order parameters of the $(r-1)$-state ABF model in regime III
\cite{ABF}.
Hence computing the sum (\ref{confsums}) amounts to computing
the order parameters of the ABF model. The fact that (finitized)
Rogers--Ramanujan identities arise from calculating order parameters
of solvable lattice models is in fact not new, and indeed the sum
(\ref{HHM}) is exactly the one encountered by Baxter in his
solution of the hard hexagon model in regime~I~\cite{Baxter81}.
The remainder of this paper is organized as follows.
In the next section we describe Melzer's polynomial identities,
their limiting Rogers--Ramanujan type form and some other
Virasoro character identities that follow from the proof of
Melzer's identities and the application of the Andrews--Bailey
construction~\cite{Bailey47,Bailey49,Andrews84}.
Then, in section~3, we compute the configuration sums $(\ref{confsums})$
using the technique developed in part I of this paper \cite{W}.
This amounts to reinterpreting the sum (\ref{confsums}) as the
grand canonical partition function of a one-dimensional gas of
charged particles obeying certain Fermi-type exclusion rules.
In section~4 we describe the original approach of ABF for computing
(\ref{confsums}) using recurrence relations.
Together with the result of section~3 this proves Melzer's
polynomial identities.
We finally end with a discussion of our result and an outlook to
related problems and generalizations.
To end this introduction we make some further remarks
on the problem described in this paper.
First, as mentioned before, an altogether different kind
of proof of Melzer's identities has recently been given for
the case of $\chi^{(r-1,r)}_{b,1}(q)$ by Berkovich \cite{Berkovich}.
This method of proof, which in fact is applicable to all
unitary minimal characters \cite{BM}, is based on recursive
instead of combinatorial arguments.\footnote{Berkovich
has subsequently proven Melzer's identities for all characters,
but his results remain unpublished \cite{Berkovich2}.}
Second, in their solution of the ABF model, Andrews, Baxter
and Forrester also considered the configuration sums $X_L(a,b,c)$,
with $c=b-1$. Hence to completely compute all configuration sums
of the ABF model, more general sums than those defined in
(\ref{confsums}) have to be considered. However, from simple
symmetry arguments \cite{ABF,Melzer} (see also section 3) one can
easily deduce that computing (\ref{confsums}) suffices to obtain
expressions for all $X_L(a,b,c)$.
Finally we remark that Melzer~\cite{Melzer} and
Kedem {\em et al.}~\cite{KKMMa} conjecture
(in the general case) four fermionic expressions for each
(finitized) character.
In this paper we give detailed proof of only two of the four.
For the remaining two representations we did not succeed
in finding a derivation in terms of a Fermi lattice-gas.
\nsection{Melzer's polynomial identities and related
Rogers--Ra\-manujan identities}
In this section we give a summary of identities proven by the
calculations carried out in the sections~\ref{sec3} and~\ref{sec4}.
First we describe the polynomial identities conjectured by
Melzer~\cite{Melzer},
and their limiting Rogers--Ramanujan type form as discovered by the
Stony Brook group~\cite{KKMMa}. Then we list two classes of character
identities for non-unitary minimal models which, as recently pointed
out by Foda and Quano \cite{FQ}, arise from Melzer's identities
and the Andrews--Bailey construction~\cite{Bailey47,Bailey49,Andrews84}.
\subsection{Identities for the (finitized) Virasoro characters
$\chi^{(r-1,r)}_{b,a}(q)$}
Before we state the polynomial identities as conjectured by Melzer,
we need some notation.
We denote the incidence matrix of the A$_{r-3}$ Dynkin diagram by
$\cal I$, with ${\cal I}_{j,k}=\delta_{j,k-1}+\delta_{j,k+1}$,
$j,k=1,\ldots,r-3$. The Cartan matrix of A$_{r-3}$ is denoted
as $C$, and is related to $\cal I$ by $C_{j,k}=2\delta_{j,k}
-{\cal I}_{j,k}$. We also define the $(r-3)$-dimensional
(column) vectors $\vec{m}$ and $\vec{\mbox{e}}_j$, $j=1,\ldots,r-3$,
by $(\vec{m})_j=m_j$ and $(\vec{\mbox{e}}_j)_k=\delta_{j,k}$, and set
$m_0=m_{r-2}=0$, $\vec{\mbox{e}}_0=\vec{\mbox{e}}_{r-2}= \vec{0}$.
With this notation, using the Gaussian
polynomials as defined in (\ref{qpoly}), Melzer's conjectures
can be stated as the following identities for
$a=1,\ldots,r-1$, $b=1,\ldots,r-2$ and
$L-|a-b|\in 2\:\mbox{\sf Z} \hspace{-0.82em} \mbox{\sf Z}\,_{\geq 0}$:\footnote{Throughout this paper
we use the notation $x \equiv y$ to mean $x \equiv y \mod{2}$
Also, the sums $\sum_{\vec{x} \equiv \vec{y}}$ and $\sum_{\vec{x}}$
are shorthand notations for $\prod_j \sum_{x_j \geq 0; \;
x_j \equiv y_j}$ and $\prod_j \sum_{x_j \geq 0}$,
respectively.}
\begin{eqnarray}
\lefteqn{
f_{a,b}
\sum_{\vec{m} \equiv \vec{Q}_{a,b}}
q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + L \, \vec{\mbox{e}}_1 +
\vec{\mbox{e}}_{r-a-1}
+\vec{\mbox{e}}_{r-b-1})_j}{m_j} } \nonumber \\
& & =
\sum_{j=-\infty}^{\infty} \left\{
q^{j\big(r(r-1)j+rb-(r-1)a\big)}
\Mult{L}{\scriptstyle \frac{1}{2}(L+a-b)-rj}
-q^{\big((r-1)j+b\big)\big(rj+a\big)}
\Mult{L}{\scriptstyle \frac{1}{2}(L-a-b)-rj} \right\},
\label{Mid}
\end{eqnarray}
with $f_{a,b}= q^{-(a-b)(a-b-1)/4}$ and
\begin{equation}
\vec{Q}_{a,b} = \vec{Q}_{a,b}^{(r-3)} =
(\vec{\mbox{e}}_{r-a-2}+\vec{\mbox{e}}_{r-a-4}+\ldots)
+(\vec{\mbox{e}}_{r-b-2}+\vec{\mbox{e}}_{r-b-4}+\ldots ).
\label{Qrest}
\end{equation}
We note that in our derivation of the left-hand side of
(\ref{Mid}) in section~3, this restriction naturally arises
in the following form, $\mod{2}$-equivalent to (\ref{Qrest}):
\begin{equation}
(\vec{Q}_{a,b})_j = \min(a-1,r-j-2)+\min(b-1,r-j-2).
\end{equation}
In ref.~\cite{Melzer}, yet another expression for the left-hand side of
(\ref{Mid}) was conjectured as
\begin{equation}
f_{a,b} \sum_{\vec{m} \equiv \vec{R}_{a,b}}
q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\frac{1}{2} m_{a-1} }
\prod_{j=1}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + L \, \vec{\mbox{e}}_1 +
\vec{\mbox{e}}_{a-1} +\vec{\mbox{e}}_{r-b-1})_j}{m_j}
\label{Mid2}
\end{equation}
where
\begin{equation}
\vec{R}_{a,b} = (r-a-1)\vec{\rho}+(\vec{\mbox{e}}_{a}+\vec{\mbox{e}}_{a+2}+\ldots)
+(\vec{\mbox{e}}_{r-b-2}+\vec{\mbox{e}}_{r-b-4}+\ldots ),
\end{equation}
with $\vec{\rho}=\sum_{j=1}^{r-3} \vec{\mbox{e}}_j$.
Clearly, for $a=1$ and for $a=r-1$ the fermionic expressions in
(\ref{Mid}) and (\ref{Mid2}) coincide.
As mentioned in the introduction,
we have no explanation of this alternative fermionic
form in terms of a Fermi-gas, and (\ref{Mid2}) is listed
only for completeness.
Taking the finitization parameter $L$ to infinity, (\ref{Mid})
leads to Rogers--Ramanujan type identities for unitary minimal
Virasoro characters. Hereto we recall the well-known
Rocha-Caridi expression for all (normalized)
characters $\chi^{(p,p')}_{r,s}(q)$ of minimal CFT $M(p,p')$,
\begin{equation}
\chi^{(p,p')}_{r,s}(q)=\frac{1}{(q)_{\infty}}
\sum_{j=-\infty}^{\infty} \left\{
q^{j(pp'j+p'r-ps)}-q^{(jp+r)(jp'+s)} \right\},
\label{RC}
\end{equation}
for $r=1,\ldots,p-1$, $s=1,\ldots,p'-1$, with $p$ and $p'$ coprime.
We thus find that the right-hand side of (\ref{Mid}) gives the
{\em bosonic} Rocha-Caridi expression for $\chi^{(r-1,r)}_{b,a}(q)$,
whereas the left-hand side leads to a {\em fermionic} counterpart,
\begin{equation}
\chi^{(r-1,r)}_{b,a}(q) = f_{a,b}
\sum_{\vec{m} \equiv \vec{Q}_{a,b} }
\frac{q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} m_{r-a-1} } }{(q)_{m_1}}
\prod_{j=2}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} +
\vec{\mbox{e}}_{r-a-1} +\vec{\mbox{e}}_{r-b-1})_j}{m_j} .
\label{SB}
\end{equation}
This result is one of the many celebrated conjectures
for fermionic character representations
made by the Stony Brook group, see e.g., refs.~\cite{KKMMa,KKMMb,DKKMM}.
An obvious symmetry of (\ref{RC}) is $\chi^{(p,p')}_{r,s}(q)=
\chi^{(p,p')}_{p-r,p'-s}(q)$. Making the transformation
$a\to r-a$ and $b\to r-b-1$ in the fermionic expression
(\ref{SB}) this symmetry is not at all manifest, except for $b=1$ and
$a=1,r-2$.
Hence we have two different fermionic representations
for each character of the unitary minimal series.
To end our discussion on Melzer's polynomial identities, we remark
that in ref.~\cite{Melzer} identities were also given for
finitizations of the characters $\chi^{(r-1,r)}_{b,a}(q)$,
with finitization parameter $L$ such that $L+a-b \not\equiv 0$.
Since these can simply be obtained from (\ref{Mid}) and (\ref{Mid2})
by the above-mentioned symmetry transformation, they are not listed
here as separate identities.
\subsection{Rogers--Ramanujan identities for
$\chi_{a,(k+1)b}^{(r,(k+1)r-1)}(q)$ and
$\chi_{b,(k+1)a}^{(r-1,(k+1)r-k)}(q)$}
It was recently pointed out by Foda and Quano~\cite{FQ},
that many new Virasoro
character identities can be obtained by applying some powerful lemmas,
proven by Bailey and Andrews, to Melzer's polynomial identities.
The main idea of these lemmas is to proof the more complicated
Rogers--Ramanujan type identities by showing that they are
a consequence of easier to proof identities.
Here we will not state the relevant lemmas but refer
the interested reader
to the work of Foda and Quano~\cite{FQ} and to the original work
of Bailey~\cite{Bailey47,Bailey49} and Andrews~\cite{Andrews84}.
In both series of Virasoro character identities given below,
we encounter the $k$ by $k$ matrix $B$ with entries
$B_{j,\ell}=\mbox{min}(j,\ell)$.
We note that this matrix is the inverse of the Cartan-type matrix
of the tadpole graph with $k$ nodes;
$(B^{-1})_{j,\ell} = 2\delta_{j,\ell}-{\cal I}_{j,\ell}^{(k)}$,
with incidence matrix of the tadpole graph given by
${\cal I}_{j,\ell}^{(k)}=\delta_{j,\ell-1}
+\delta_{j,\ell+1}+\delta_{j,\ell}\,\delta_{j,k}$, $j,\ell=1,\ldots,k$.
We will also use the $k$-dimensional vectors $\vec{n}$ and
$\vec{\varepsilon}_k$, whose $j$-th entries read $n_j$ and $\delta_{k,j}$,
respectively.
\subsubsection{$\chi_{a,(k+1)b}^{(r,(k+1)r-1)}(q)$}
Substituting the {\em Bailey pair} read off from (\ref{Mid}) into the
{\em Bailey chain} of length $k$, we obtain
\begin{eqnarray}
\chi_{a,(k+1)b}^{(r,(k+1)r-1)}(q)
&\stackrel{\vphantom{\case{1}{2}} (a\equiv b)}{=}&
f_{a,b} \; q^{-k(a-b)^2/4}
\sum_{\vec{n}}
\sum_{\vec{m} \equiv \vec{Q}_{a,b}}
\frac{q^{\,\vec{n}^T B \: \vec{n}}}
{(q)_{n_1} \ldots (q)_{n_{k-1}} (q)_{2n_k}} \nonumber \\
& & \times \;
q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mults{\case{1}{2}({\cal I} \, \vec{m} + 2n_k \, \vec{\mbox{\scriptsize e}}_1 +
\vec{\mbox{\scriptsize e}}_{r-a-1}
+\vec{\mbox{\scriptsize e}}_{r-b-1})_j}{m_j}
\nonumber \\
&\stackrel{\vphantom{\case{1}{2}} (a\not\equiv b)}{=}&
f_{a,b} \; q^{-k\bigl((a-b)^2-1\bigr)/4}
\sum_{\vec{n}}
\sum_{\vec{m} \equiv \vec{Q}_{a,b} }
\frac{q^{\,\vec{n}^T B \: (\vec{n} +\vec{\varepsilon}_k)}}
{(q)_{n_1} \ldots (q)_{n_{k-1}} (q)_{2n_k+1}}
\label{AB} \\
& & \times \;
q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mults{\case{1}{2}({\cal I} \, \vec{m} + (2n_k+1) \, \vec{\mbox{\scriptsize e}}_1 +
\vec{\mbox{\scriptsize e}}_{r-a-1}
+\vec{\mbox{\scriptsize e}}_{r-b-1})_j}{m_j}
\nonumber
\end{eqnarray}
valid for all $k\geq 1$, $a=1,\ldots,r-1$, $b=1,\ldots,r-2$.
The proof of this result for $a=1$ was first noted by Foda and
Quano \cite{FQ}, using the proof of Melzer's identities for $a=1$ as
established by Berkovich~\cite{Berkovich}. The fermionic expression
in (\ref{AB}) can also be found in ref.~\cite{BM}.
\subsubsection{$\chi_{b,(k+1)a}^{(r-1,(k+1)r-k)}(q)$}
Substitute the {\em dual} Bailey pair
obtained from (\ref{Mid}) into the
Bailey chain of length $k+1$. Then make the change
of variables $m_j\to m_{j+1}$, followed by $2 n_{k+1}+|a-b|\to m_1$,
$n_k \to n_k +\frac{1}{2}(m_1-|a-b|)$ and $r\to r-1$.
Finally, interchanging $a$ and $b$ then using
\begin{equation}
\left( \vec{Q}_{a,b}^{(r-3)} \right)_j \equiv \left\{
\begin{array}{ll}
\left( \vec{Q}_{b,a}^{(r-4)} \right)_{j-1}
\quad & j=2,\ldots,r-3 \\
a-b & j=1,
\end{array} \right.
\end{equation}
true for $a=1,\ldots,r-3$, $b=1,\ldots,r-2$,
yields
\begin{eqnarray}
\chi_{b,(k+1)a}^{(r-1,(k+1)r-k)}(q) &=&
f_{a,b} \; q^{-k(a-b)^2/4}
\sum_{\vec{n} }
\sum_{\vec{m} \equiv \vec{Q}_{a,b} }
\frac{q^{ (\vec{n} + \case{1}{2} m_1 \vec{\varepsilon}_k)^T B \:
(\vec{n} + \case{1}{2} m_1 \vec{\varepsilon}_k)}}
{(q)_{n_1} \ldots (q)_{n_k}}
\nonumber \\
& & \times \;
\frac{q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} m_{r-a-1} }}{(q)_{m_1}}
\prod_{j=2}^{r-3}
\Mults{\case{1}{2}({\cal I} \, \vec{m} +
\vec{\mbox{\scriptsize e}}_{r-a-1} +\vec{\mbox{\scriptsize e}}_{r-b-1})_j}{m_j},
\label{dualB}
\end{eqnarray}
valid for all $k\geq 0$, $a=1,\ldots,r-3$, $b=1,\ldots,r-2$.
Note that for $k=0$, corresponding to a Bailey chain
of length 1, we actually recover a subset of the character
identities (\ref{SB}) for $M(r-1,r)$.
For $a=b=1$, (\ref{dualB}) was conjectured in ref.~\cite{KKMMa}.
The proof for $a=1$ can again be found in ref.~\cite{FQ},
though the actual form of the fermionic side therein
rather differs due to the sequence of the transformations
carried out above. The fermionic form (\ref{dualB})
can also be found in ref.~\cite{BM}.
\nsection{Fermionic solution of the ABF model}\label{sec3}
We now come to the main part of this paper, the
evaluation of the one-dimensional configuration
sums (\ref{confsums}) of the ABF model. This yields,
up to the prefactor $f_{a,b}$,
the left-hand side of the identity (\ref{Mid}).
To establish this, we first reformulate the sum (\ref{confsums})
as the generating function
of certain restricted lattice paths. We then compute
this generating function by identifying each path
as a configuration of charged fermions on a one-dimensional
lattice. This identification allows us to view $X_L(a,b)$
as the grand-canonical partition function
of a one-dimensional Fermi-gas.
Because of the one-dimensional nature of this gas, its
partition function can readily be computed.
\subsection{Restricted lattice paths}
To reformulate the sum (\ref{confsums}) in terms of lattice paths,
we first give some basic definitions.
\begin{definition}
An ordered sequence of spins
$\{\sigma_0,\sigma_1,\ldots,\sigma_{L+1}\}$ is called
admissible if
\begin{itemize}
\item $\sigma_j\in \{0,1,\ldots,r-2\}$ for $j=0,\ldots,L+1$,
\item $\mbox{$|\sigma_{j+1}-\sigma_j|=1$}$ for $j=0,\ldots,L$, and
\item $\sigma_0=a-1$, $\sigma_L=b-1$ and $\sigma_{L+1}=b$.
\end{itemize}
\end{definition}
\begin{definition}
Let $\{\sigma_0,\sigma_1,\ldots,\sigma_{L+1}\}$ be an admissible
sequence of spins. Plot all pairs $(j,\sigma_j)$
in the $(x,y)$-plane and interpolate
between each pair of neighbouring points by a straight line
segment. The resulting graph is called a restricted lattice path.
\end{definition}
An example of a restricted lattice path
for $a=3$ and $b=5$ is shown in Figure~\ref{fig1}.
\begin{figure}[t]
\epsfxsize = 10cm
\centerline{\epsffile{path.ps}}
\caption{An example of a restricted lattice
path in rlp$(0,r)$.}
\label{fig1}
\end{figure}
To write the one-dimensional configuration sum
as a sum over restricted lattice paths, first notice that
the restrictions on the $\sigma$'s in
(\ref{confsums}) precisely correspond to
those defining an admissible sequence of spins.
Consequently, each restricted lattice path
corresponds to one of the terms in the sum (\ref{confsums})
and, conversely, each term in the sum corresponds to
a restricted lattice path.
Given an admissible sequence, its total weight
is decomposed as follows. If $\sigma_{j-1}<\sigma_j<\sigma_{j+1}$ or
$\sigma_{j-1}>\sigma_j>\sigma_{j+1}$ this contributes a factor
$q^{j/2}$ and if
$\sigma_{j-1}<\sigma_j>\sigma_{j+1}$ or
$\sigma_{j-1}>\sigma_j<\sigma_{j+1}$ this contributes a factor 1.
In terms of the restricted lattice paths this simply means that
for each integer point $j$ along the $x$-axis we get a factor
1 if $(j,\sigma_j)$ is an extremum and a factor $q^{j/2}$
otherwise. Here the terminals of a path are to be viewed
as extrema.
Writing this in the language of statistical mechanics we get,
setting $q=\exp(-\beta)$,
\begin{equation}
X_L(a,b) = \sum_{\mbox{\scriptsize restricted lattice paths}}
\mbox{e}^{\; \displaystyle -\beta \sum_{j=1}^L E(j)},
\label{Xpaths}
\end{equation}
with energy function $E$ given by
\begin{equation}
E(j) = \left\{
\begin{array}{ll}
0 & \mbox{if the path has an extremum at ($x$-position) $j$} \\
\frac{1}{2} j
\qquad & \mbox{otherwise.}
\end{array}\right.
\label{energy}
\end{equation}
Each of the lattice paths in the sum (\ref{Xpaths})
starts in $(0,a-1)$, ends in $(L,b-1),\; (L+1,b)$ and
is restricted to the strip $0\leq y \leq r-2$.
We now define rlp$(\mu,r)$ as the set of all restricted
lattice paths with minimal $y$ value equal to $\mu$ and
maximal $y$ value less or equal to $r-2$.
Hence we can write
\begin{equation}
X_L(a,b)=\sum_{\mu=0}^{\min(a,b)-1} \Xi_L(a,b;\mu,r),
\end{equation}
with
\begin{equation}
\Xi_L(a,b;\mu,r) = \sum_{\mbox{\scriptsize rlp}(\mu,r)}
\mbox{e}^{\; \displaystyle -\beta \sum_{j=1}^L E(j)}.
\label{XiLabmu}
\end{equation}
Noting the obvious relation
$\Xi_L(a,b;\mu,r) = \Xi_L(a-\mu,b-\mu;0,r-\mu)$
gives
\begin{equation}
X_L(a,b)=\sum_{\mu=0}^{\min(a,b)-1} \Xi_L(a-\mu,b-\mu;0,r-\mu),
\label{summu}
\end{equation}
and we conclude that to compute $X_L(a,b)$ it suffices to compute
sum (\ref{XiLabmu}) for $\mu=0$, and arbitrary $a$, $b$ and $r$.
So far we only have reformulated the problem of
computing $X_L(a,b)$, and it is by no means clear that
$\Xi_L(a,b):=\Xi_L(a,b;0,r)$ is any simpler to evaluate than
(\ref{confsums}). To make some real progress,
we will show in the next section that $\Xi_L(a,b)$ can be
viewed as the grand canonical partition function
of a one-dimensional gas of charged fermions.
In other words, each path in rlp$(0,r)$
can be viewed as a configuration of an appropriately
defined Fermi-gas.
Now decomposing the sum over all Fermi-gas configurations
into a sum over configuration with fixed particle
content (FC) and a sum over the particle content (C),
we get
\begin{equation}
\Xi_L(a,b) = \sum_{\mbox{\scriptsize C}}
Z(\mbox{C};a,b),
\label{ZCab}
\end{equation}
with $Z(\mbox{C};a,b)$
the partition function of the 1-dimensional
Fermi-gas,
\begin{equation}
Z(\mbox{C};a,b) = \sum_{\mbox{\scriptsize FC}}
\mbox{e}^{\; \displaystyle -\beta \sum_{j=1}^L E(j)}.
\end{equation}
\subsection{A one-dimensional Fermi-gas}
To interpret each restricted lattice path in rlp$(0,r)$
as a configuration of particles, we need some more terminology.
In fact, since some of the concepts introduced below are
somewhat awkward to describe, but easily explained pictorially,
we state some definitions purely graphically.
In the previous section restricted lattice path were
introduced as path from $(0,a-1)$ to $(L,b-1)$, $(L+1,b)$,
restricted to the strip $0\leq y \leq r-2$, such that
$y_{j+1}-y_j=\pm 1$ for all consecutive points
$(j,y_j)$ and $(j+1,y_{j+1})$ on the path.
We somewhat relax these conditions by defining a
{\em lattice path} as
\begin{definition}
A lattice path is a restricted lattice path with
arbitrary (integer) begin- and endpoint.
\end{definition}
In particular, if a lattice path ends in $(j,y_j)$, the
$y$-coordinate of the second-last point can either be
$y_j-1$ or $y_j+1$.
We use the previous definition to define a very important
object, {\em a complex}.~\footnote{In ref.~\cite{Bressoud},
Bressoud has given a lattice path interpretation of the Andrews--Gordon
generalizations of the Rogers--Ramanujan
identities~\cite{Gordon,Andrews74}. In Bressoud's terminology
a complex corresponds to a {\em mountain}.}
This will be used subsequently to decompose each
restricted lattice path into particles.
\begin{definition}
A bulk complex is a lattice path from $(j,y_j)$ to $(k,y_k)$,
with $(j,y_j)$ and $(k,y_k)$ connected by
a dashed horizontal line,
such that $y_j=y_k$, $y_{\ell}>y_j$ for all $j<\ell<k$.
A left-boundary complex is a lattice path
from $(0,a-1)$ to $(j,0)$, such that
$y_k>0$ for all $k<j$, and with
$(0,0)$ and $(j,0)$ connected by a horizontal dashed line
and $(0,0)$ and $(0,a-1)$ connected by a vertical
solid line.
A right-boundary complex is a lattice path
from $(j,0)$ to $(L,b-1)$, $(L+1,b)$,
such that $y_k>0$ for all $k>j$ and with
$(L+1,0)$ and $(j,0)$ connected by a horizontal dashed line
and $(L+1,0)$ and $(L+1,b)$ connected by a vertical
solid line.
\end{definition}
Examples of a left-boundary, bulk and
right-boundary complex can be found in Fig.~\ref{complex}.
\begin{figure}[t]
\epsfysize = 3.5cm
\centerline{\epsffile{complex.ps}}
\caption{Typical examples of a left-boundary, bulk, and
right-boundary complex.}
\label{complex}
\end{figure}
With respect to the above definition we remark that
the term complex is chosen since we
wish to view each complex as a collection of charged
particles moved on top of each other.
To make this explicit, we define particles
in the following two definitions.
\begin{definition}
A \underline{pure}
bulk particle of charge $j$ is a bulk complex with a single local
maximum of height $j$ (measured with respect to its dashed line).
A \underline{pure} left-boundary particle of charge $(a-1)/2$
is a left-boundary complex with a single local maximum,
located at $(0,a-1)$.
A \underline{pure} right-boundary particle of charge $b/2$
is a right-boundary complex with a single local maximum.
\end{definition}
The graphical representation of pure particles is given in
Fig.~\ref{pure}.
\begin{figure}[t]
\epsfysize = 3.5cm
\centerline{\epsffile{pure.ps}}
\caption{The graphical representation of pure particles.
The charges are, from left to right,
$(a-1)/2$, $j$ and $b/2$, respectively.}
\label{pure}
\end{figure}
To introduce the more general idea of a particle, we need some
simple terminology.
\begin{itemize}
\item The {\em peak} of a bulk complex is the
left-most highest point. Similarly,
the peak of a particle is its highest point.
\item The {\em origin} of a particle or complex
is the left- and down-most point. \\
The {\em endpoint} of a particle or complex
is the right- and down-most point. \\
The {\em baseline} of a particle or complex
is the dashed line connecting the begin and
endpoint.
\item The {\em contour} of a particle or complex
is its part drawn with solid lines.
\end{itemize}
Using this we define
\begin{definition}
A bulk particle of charge $j$ is a pure bulk particle of charge $j$,
whose contour is interrupted at arbitrary integer points by horizontal
dashed lines of even length.
A left-boundary particle of charge $(a-1)/2$ is a pure
left-boundary particle of charge $(a-1)/2$, whose contour to the right
of $(0,a-1)$ is interupted at arbitrary integer points by horizontal
dashed lines of even length.
A right-boundary particle of charge $b/2$ is a pure
right-boundary particle of charge $b/2$, whose contour to the left
of $(L,b-1)$ is interupted at arbitrary integer points by horizontal
dashed lines of even length.
\end{definition}
Typical examples of particles are shown in Fig.~\ref{unpure}.
We note that for later convenience the contour of the boundary
particles is drawn with thicker lines than that
of the bulk particles.
\begin{figure}[bt]
\epsfysize = 3.5cm
\centerline{\epsffile{unpure.ps}}
\caption{
Typical examples of a left-boundary, bulk
and right-boundary particle.
The charges of the particles are
$(a-1)/2$, $j$ and $b/2$, respectively.}
\label{unpure}
\end{figure}
With the above set of definitions we now give a prescription
to divide each restricted lattice path into particles.
This will be done by giving an algorithm that divides a complex
into a particle and several smaller complexes.
Each of these new complexes is either a particle
or is again divided into a particle
and yet smaller complexes.
This procedure is continued until the entire complex
is divided into particles. Since each lattice path can trivially
be divided into complexes, this gives a procedure to
divide any restricted lattice path into particles.
\begin{description}
\item[(0)]
Draw a dashed line along the $x$-axis from $(0,0)$ to $(L+1,0)$,
and draw bold lines from $(0,0)$ to $(0,a-1)$ and $(L+1,0)$
to $(L+1,b)$.
This divides each restricted lattice path into a
left-boundary complex, a right-boundary complex and a number of
bulk complexes.
For the restricted lattice path of Fig.~\ref{fig1}, we for
example get 4 complexes, 2 of which are of bulk-type.
If $a=1$, the left-boundary complex is absent.
Now consider each of the complexes obtained above. If such
a complex is a particle (in which case it is pure),
we are done with it.
If not, go to step (1) in case of a bulk complex and to
(1$_L$) and (1$_R$) in case of a left- and
right-boundary complex, respectively.
\item[(1)]
Start at the peak of the complex and
move down to the right along the contour
till the endpoint of the complex.
When a local minimum is reached, i.e., the
contour starts going up again, we draw a dashed line
from this local minimum to the right until we cross the contour.
At that point we move further down along the contour.
If another minimum occurs we repeat the above, et cetera.
Repeat the above now moving to the left. That is,
start from the peak of the complex
and move down to the left till the origin of the complex.
If a local minimum is reached we
draw a dashed line to the left and continue our movement down when
the dashed line intersects the contour.
As a result of the above step we have divided the complex
into a particle (which is not pure)
and several (at least one) smaller complexes.
The peak and the baseline of the particle are the peak and
the baseline of the original complex.
Now go to (2).
\item[(1$_L$)]
Start from $(0,a-1)$.
Move to the right of this point down along the contour of the
complex till its endpoint.
If a local minimum is reached (which could be the point
$(0,a-1)$ itself), draw a dashed line
from this minimum to the right, until the contour is crossed.
At that point move further down along the
contour. If another minimum occurs we repeat the above, et cetera.
As a result of the above step we have divided the
left-boundary complex into a left boundary particle
and several (at least one) smaller bulk complexes.
To treat these smaller bulk complexes, go to (2).
\item[(1$_R$)]
Start from $(L+1,b)$.
Move to the left of this point down along the contour of the
complex till its endpoint.
If a local minimum is reached,
draw a dashed line from this minimum
to the left until the contour is crossed.
At that point move further down along the
contour. If another minimum occurs repeat the above, et cetera.
As a result of the above step we have divided the
right-boundary complex into a right-boundary particle
and several (at least one) smaller bulk complexes.
To treat these smaller bulk complexes, go to (2).
\item[(2)]
Scan each of the smaller bulk complexes. If such a
complex is a bulk particle (in which case it is pure),
we are done with it.
If not repeat step (1) for this complex.
\end{description}
We note that the above procedure converges, since the number of
local maxima of a restricted lattice path is finite.
In Fig.~\ref{content}, we have carried out the procedure
for the restricted lattice path of Fig.~\ref{fig1},
thereby identifying the corresponding configuration of
particles.\footnote{After having identified all particles, we
implicitly assume the step of (re)drawing the contour of the
boundary particles with fat lines.}
\begin{figure}[t]
\epsfxsize = 10cm
\centerline{\epsffile{content.ps}}
\caption{The particle configuration corresponding to the
restricted lattice path of Fig.~1.}
\label{content}
\end{figure}
Thanks to the above algorithm, each restricted lattice path
in rlp$(0,r)$ can now be viewed as a particle configuration.
In particular, since the maximal height of a path is
$r-2$, we have bulk particles of charge 1 up to $r-2$, as well
as a left-boundary particle of charge $(a-1)/2$ and a
right-boundary particle of charge $b/2$.
The contour of a bulk particle of charge $j$ consists
of $j$ up and $j$ down steps, the contour of a
left-boundary particle of $a-1$
down steps and the contour of a
right-boundary particle of $b$
up steps.
Letting $n_j$ denote the number of bulk particles
of charge $j$, we thus have the completeness relation
\begin{equation}
a + b - 1 + 2\sum_{j=1}^{r-2} j\, n_j = L+1.
\label{completeness}
\end{equation}
Using this relation,
$n_{r-2}$ can be computed given the occupation numbers
$n_1,\ldots,n_{r-3}$. For this reason (and anticipating
things to come), we define the column
vector $\vec{n}= \,^T(n_1,\ldots,n_{r-3})$,
and when we say
``a restricted lattice path has particle content $C=\vec{n}\,$'',
we mean by this the particle content $C=\{n_1,\ldots,n_{r-2}\}$
subject to the restriction (\ref{completeness}).
Having associated a configuration of
particles with each path in rlp$(0,r)$,
we define rlp$(\vec{n})$ as the subset of
paths in rlp$(0,r)$, with particle content
$\vec{n}$.
This puts us in a position to properly define what
we mean by the Fermi-gas partition function as
introduced in (\ref{ZCab}),
\begin{equation}
Z(C;a,b) = Z(\vec{n};a,b) =
\sum_{\mbox{\scriptsize rlp}(\vec{n})}
\mbox{e}^{\; \displaystyle -\beta \sum_{j=1}^L E(j)},
\end{equation}
with energy function defined in (\ref{energy}).
So far, we have repeatedly used the term
Fermi-gas, without any clear motivation.
Clearly, we have defined all allowed
configurations of our one-dimensional system of charged
particles, as well its Hamiltonian or energy
function, but the actual nature of the system remains
rather elusive.
However, in our actual computation of $Z$, in the next
subsection, it turns out to be expedient to
define rules of motion that allow one to
obtain any configuration with content $\vec{n}$
from a given so-called minimal configuration
with the same content.
These rules of motion have a clear fermionic character, in
that particles of the same charge cannot exchange
position, unlike particles of different charge.
\subsection{Computation of $Z(\vec{n};a,b)$.}
In this section we compute the partition function
of the one-dimensional Fermi-gas.
Throughout the section we assume the
particle content to be $\vec{n}$.
To compute the sum over all particle configurations,
we first select a particular configuration called
the {\em minimal configuration}.\footnote{From a statistical
mechanics point of view {\em ground state configuration}
may be more appropriate, but we prefer
to conform to our earlier naming in ref.~\cite{W}.}
It will be defined purely graphically.
\begin{definition}
The configuration shown in Fig.~\ref{min} is called the
minimal configuration. Here each bulk particle of charge
$j$ should be repeated $n_j$ times, i.e.,
$$
\centerline{\epsffile{mult.ps}}
$$
\end{definition}
Note that in the minimal configuration
\begin{itemize}
\item All (bulk) particles are positioned as much to the right and up
as possible, the baseline of the particles of charge $j$ having
$y$-coordinate equal to $\min(b-1,r-j-2)$.
\item The particles are positioned
in order of decreasing charge.
\item Apart from the right-boundary particle, all particles are pure.
\end{itemize}
\begin{figure}[t]
\epsfxsize = 11cm
\centerline{\epsffile{min.ps}}
\caption{The restricted lattice path corresponding to the
minimal configuration of particles.
Here each bulk particle of charge $j$
has to be copied $n_j$ times. The dashed lines are the
baselines of the particles.}
\label{min}
\end{figure}
\subsubsection{Contribution of the minimal configuration}
To compute the weight of the minimal configuration, we use that the
energy $E_j(x)$ of a pure bulk particle
of charge $j$, with origin at position $x$
and endpoint at position $x+2j$,
is given by
\begin{equation}
E_j(x)=
\frac{1}{2} \; \sum_{
{k=1 \atop k\neq j}}^{2j-1} (k+x)
= (j+x)(j-1).
\label{Ejx}
\end{equation}
Similarly, we get for the energy $E_a$ of the pure
left-boundary particle
with charge $(a-1)/2$,
\begin{equation}
E_a=\frac{1}{4}(a-1)(a-2).
\label{Ea}
\end{equation}
A bit more work is required to obtain the energy $E_b$ of the
right-boundary particle with charge $b$, since its contour
is broken into $b$ segments all of length 1.
Summing up the $b$ different contributions leads to
\begin{equation}
E_b=\frac{1}{4}(b-1)(2a+b-2) + \sum_{j=r-b}^{r-2}
j(b-r+j+1) n_j.
\end{equation}
Using the above three results, we compute the energy of
the minimal configuration as
\begin{eqnarray}
E_{\min}
&=& E_a+E_b+\sum_{j=1}^{r-2} \sum_{\ell=1}^{n_j}
E_j\biggl(a-2+\min(b,r-j-1)+2j(\ell-1)
+ 2\sum_{k=j+1}^{r-2} k \, n_k \biggr)
\nonumber \\
&=& \sum_{j=1}^{r-2} (j-1) n_j \biggl( j \, n_j +
2 \sum_{k=j+1}^{r-2} k \, n_k \biggr) +
(a+b-2) \sum_{j=1}^{r-2}(j-1) n_j \\
& & + \sum_{j=r-b}^{r-2} (b-r+j+1) n_j
+ \frac{1}{4}(a-1)(a-2) + \frac{1}{4}(b-1)(2a+b-2).
\nonumber
\end{eqnarray}
To simplify this expression, we eliminate $n_{r-3}$ using the
completeness relation (\ref{completeness}). This yields
\begin{eqnarray}
\lefteqn{
E_{\min} =
\sum_{j=1}^{r-3}
\left(
\sum_{k=1}^j \frac{k (r-j-2)}{r-2} +
\sum_{k=j+1}^{r-3} \frac{j (r-k-2)}{r-2}
\right)
n_j \, n_k }
\nonumber \\
& & \qquad
- \left(
\sum_{j=1}^{r-b-1} \frac{j (b-1)}{r-2} +
\sum_{j=r-b}^{r-3} \frac{(r-b-1)(r-j-2)}{r-2}
+ L \sum_{j=1}^{r-3} \frac{r-j-2}{r-2} \right) n_j \\
& & \qquad
+ \: \frac{L^2(r-3) + 2 L (b-1) -(a-1)(r-a-1)+(b-1)(r-b-1)}{4(r-2)}.
\nonumber
\end{eqnarray}
We now recall the definition of the inverse Cartan matrix of
the Lie algebra A$_{r-3}$,
\begin{equation}
\renewcommand{\arraystretch}{2.4}
C^{-1}_{j,k} =
\left\{
\begin{array}{cc}
\displaystyle
\frac{k(r-j-2)}{r-2} \qquad & k \leq j \\
\displaystyle
\frac{j(r-k-2)}{r-2} & k \geq j .
\end{array}
\right.
\label{invC}
\end{equation}
Using this, we finally obtain
\begin{lemma}\label{lemmin}
The energy of the minimal configuration is given by
\begin{eqnarray}
\lefteqn{
E_{\min}
=\sum_{j,k=1}^{r-3}
\left(n_j-\frac{L}{2} \, \delta_{j,1}-\frac{1}{2}\, \delta_{j,r-b-1}
-\frac{1}{2}\, \delta_{j,r-a-1} \right)
C^{-1}_{j,k} } \nonumber \\
& & \qquad \qquad\qquad \qquad \qquad \times
\left(n_k-\frac{L}{2}\, \delta_{k,1}-\frac{1}{2}\, \delta_{k,r-b-1}
+\frac{1}{2}\, \delta_{k,r-a-1} \right).
\label{Emin}
\end{eqnarray}
\end{lemma}
\subsubsection{Contribution of the non-minimal configurations}
To compute the contribution to the partition function of the
other configurations, we define rules of motion
which generate all non-minimal configurations from the minimal one.
These rules break up into several different {\em
elementary moves} as follows.
\begin{definition}\label{defmoves}
Let $X=\{(x_1,y_1),(x_2,y_2),(x_3,y_3),(x_4,y_4)\}$
denote a sequence of four points on the contour of a
configuration, each pair of consecutive points connected
straight lines, such that the contour
in between $(x_1,y_1)$ and $(x_4,y_4)$ does not
belong to a boundary particle.
We may then replace this sequence by a new sequence
of four points as follows.
\begin{description}
\item[move $L_u$:]
If $y_4\leq y_2 <y_3<y_1$,
\noindent
$\qquad \qquad \qquad \quad
L_u(X)=
\{(x_1,y_1),(x_2-1,y_2+1),(x_3-1,y_3+1),(x_4,y_4)\}$.
\item[move $R_d$:]
If $y_4<y_2<y_3\leq y_1$,
\noindent
$\qquad \qquad \qquad \quad
R_d(X)=
\{(x_1,y_1),(x_2+1,y_2-1),(x_3+1,y_3-1),(x_4,y_4) \}$.
\item[move $L_d$:]
If $y_1< y_3 <y_2\leq y_4$,
\noindent
$\qquad \qquad \qquad \quad
L_d(X)=
\{(x_1,y_1),(x_2-1,y_2-1),(x_3-1,y_3-1),(x_4,y_4)\}$.
\item[move $R_u$:]
If $y_1\leq y_3 <y_2< y_4$,
\noindent
$\qquad \qquad \qquad \quad
R_u(X)=
\{(x_1,y_1),(x_2+1,y_2+1),(x_3+1,y_3+1),(x_4,y_4) \}$.
\end{description}
\end{definition}
Besides these ``bulk-type'' moves we need some special
boundary moves.
\begin{definition}\label{defmovesl}
Let $X=\{(x_1,y_1),(x_2,y_2),(x_3,y_3),(x_4,y_4)\}$
be four points on the contour of a configuration,
each pair of consecutive points connected
by a straight line.
We may then replace $X$ as follows.
\begin{description}
\item[move $L'_u$:]
Let $(x_1,y_1)=(x_2-1,y_2+1)$. If $y_2=y_4<r-2$ and
the contour between the first two points belongs to the
left-boundary particle,
\noindent
$\qquad \qquad
L'_u(X)=
\{(x_2-1,y_2+1),(x_3-1,y_3+1),(x_4-1,y_4+1),(x_4,y_4)\}$,
\noindent
where the contour between the last two points belongs to the
left-boundary particle.
\item[move $R'_d$:]
Let $(x_4,y_4)=(x_3+1,y_3-1)$. If $y_1=y_3<2$
and the contour between the last two points belongs to the
left-boundary particle,
\noindent
$\qquad \qquad
R'_d(X)=
\{(x_1,y_1),(x_1+1,y_1-1),(x_2+1,y_2-1),(x_3+1,y_3-1)\}$,
\noindent
where the contour between the last two points belongs to the
left-boundary particle.
\item[move $L'_d$:]
Let $(x_1,y_1)=(x_2-1,y_2-1)$. If $y_2=y_4<y_3$
and the contour between the first two points belongs to the
right-boundary particle,
\noindent
$\qquad \qquad
L'_d(X)=
\{(x_2-1,y_2-1),(x_3-1,y_3-1),(x_4-1,y_4-1),(x_4,y_4)\}$,
\noindent
where the contour between the last two points belongs to the
right-boundary particle.
\item[move $R'_u$:]
Let $(x_4,y_4)=(x_3+1,y_3+1)$. If $y_1=y_3<y_2<r-2$,
$y_3<b-1$
and the contour between the last two points belongs to the
right-boundary particle,
\noindent
$\qquad \qquad
R'_u(X)=
\{(x_1,y_1),(x_1+1,y_1+1),(x_2+,y_2+1),(x_3+1,y_3+1)\}$,
\noindent
where the contour between the first two points belongs to the
right-boundary particle.
\end{description}
\end{definition}
For the graphical interpretation of this long list of moves,
see Fig.~\ref{moves}.
\begin{figure}[hbt]
\epsfxsize = 13cm
\centerline{\epsffile{moves.ps}}
\caption{
(a) The moves $L_u$ and $R_u$.
(b) The moves $L_d$ and $R_u$.
(c) The moves $L_u'$ and $R_u'$.
(d) The moves $L_d'$ and $R_u'$.}
\label{moves}
\end{figure}
To fully appreciate these moves, we list its main characteristics in
several lemmas, which are at the core of our fermionic computation
of the one-dimensional configuration sums.
\begin{lemma}
The elementary moves are reversible. That is,
if there is a move of type $M_s^p$ from a configuration $C$
to a configuration $C'$, then there is a move of type
$\bar{M}_{\bar{s}}^p$
from $C'$ to $C$.
Here $M=L$ or $R$, $s=u$ or $d$, $p=\quad,\, '$ or $''$ and
$\bar{R}=L$, $\bar{L}=R$, $\bar{u}=d$ and $\bar{d}=u$.
\end{lemma}
Proof: Let us show this for $L_u$. The other moves follow in
similar manner.
Let $X$ be a sequence of
four extrema as in definition~\ref{defmoves}, satisfying
$y_4\leq y_2 <y_3<y_1$. Hence we can carry out $L_u$ to obtain
$X'=L_u(X)=\{(x'_1,y'_1),(x'_2,y'_2),(x'_3,y'_3),(x'_4,y'_4)\}$.
{}From the definition of the move $L_u$,
we find that $y_4'\leq y_2'-1 < y_3'-1<y_1$. We rewrite this to obtain
$y_4'< y_2' < y_3'\leq y_1$ and hence we can carry out the
move $R_d$ to obtain $R_d(X')=X$.
$\Box$
\begin{lemma}\label{lem2}
The moves leave the particle content $\vec{n}$ fixed.
\end{lemma}
Proof: This follows immediately from the graphical representation
of the moves shown in Fig.~\ref{moves}, where the dashed lines
represent the baselines of the pure particles being moved.
Note here that the graphical representations of the
moves $R_u$ and $L_d$ are the generic cases.
Performing a move of type $R_u$ to a sequence $X$ as defined in
definition~\ref{defmoves}, with $y_2=y_4-1$, may lead to a ``jump''
of the baseline.
A similar thing may happen when performing a move of type $L_d$
to a sequence with $y_2= y_4$:
$$
\hspace{-5 cm}
\epsfxsize = 6 cm
\centerline{\epsffile{jump.ps}}
\hspace{-4 cm}
\Box$$
\begin{lemma}
Given the minimal configuration, we cannot
make any of the $R$-type moves.
\end{lemma}
Proof:
We can only make moves of type $R_d$ if we have
a sequence of $X$ as in definition~\ref{defmoves}, with
$y_4<y_2$. Clearly this does not occur.
We can only make moves of type $R_u$ if we have
a sequence $X$, with $y_2 < y_4$.
Again this does not occur.
We cannot make a move of type $R'_d$ since the
left-boundary particle is in its pure form.
Finally, we cannot make a move of type $R'_u$
since all particles of charge $j\geq r-b-1$ have their peak
at $y=r-2$, and all particles of charge $j\leq r-b-1$ have their
endpoint at $y=b-1$. $\Box$
\begin{lemma}
If a configuration is not the minimal one, we can always
make a move of type $R$.
\end{lemma}
Proof:
By construction the minimal configuration is the only configuration
that does not meet any of the conditions required for one of the
$R$-type moves. In particular, all maxima (apart from the
initial point of the path) are of decreasing order and all
minima of increasing order. This completely fixes the path.
If one of these two properties is broken somewhere along the
path, we can always make an $R$-type move. $\Box$
These first four lemmas can be combined to give the
following proposition:
\begin{proposition}
All non-minimal paths are generated by moves of type $L$
from the minimal configuration.
All non-minimal configurations can be reduced to the
minimal configuration by moves of type $R$.
\end{proposition}
Having established the above proposition, we can perform the actual
calculation of the generation function $\cal C$
of the moves of type $L$.
Again we prepare some lemmas to obtain the desired result.
\begin{lemma}\label{genqq}
Each move of type $L$ generates a factor $q$.
\end{lemma}
Proof: We show this for the typical case of move $L_u$.
The total energy $E$ of a sequence of extrema $X$ is
\begin{equation}
E=\sum_{{j=x_1+1 \atop j\neq x_2,x_3}}^{x_4-1} j.
\end{equation}
Similarly, the energy $E'$ of the sequence $X'=L_u(X)$ is
\begin{equation}
E=\sum_{{j=x_1+1 \atop j\neq x_2-1,x_3-1}}^{x_4-1} j.
\end{equation}
Hence we find
\begin{equation}
\mbox{e}^{-\beta (E'-E)} = q^{E'-E} = 1. \qquad \quad \Box
\end{equation}
In the following it will be convenient to label the
bulk particles in the minimal configuration, letting
$p_{j,\ell}$ denote the $\ell$-th particle of charge $j$,
counted from the left.
To now generate all non-minimal configurations, we
give an ordering for carrying out the moves of type $L$.
\begin{itemize}
\item
The particle $p_{j,\ell}$ is moved to the left using
moves of type $L$, prior to any of the particles
$p_{k,m}$, with $k\leq j$, and with $m>\ell$ if $k=j$.
\end{itemize}
Assuming this order (which will be justified later), we have
\begin{lemma}\label{lemmj}
The maximal number of $L$-type moves $p_{j,1}$ can make is
\begin{equation}
m_j=2 \sum_{k=j+1}^{r-2} (k-j) \, n_k
+\min(a-1,r-j-2) + \min(b-1,r-j-2).
\label{mj}
\end{equation}
\end{lemma}
Proof:
We proof this lemma in two steps. In the first step
(\ref{mj}) is shown to be true
for the minimal configuration,
and in the second step it is shown that $m_j$ is invariant under
having moved the particles $p_{k,m}$, with $k>j$, prior
to $p_{j,1}$.
Let us start to
calculate the number of $L$-type moves
needed to exchange the position of two particles of charge
$k$ and $j$, $k>j$, with $j$ positioned immediately
to the right of $k$.
In such a configuration of two particles we have a
sequence $X=\{(x_1,y_1),\ldots,(x_5,y_5)\}$ of
points connected by straight lines,
with $y_1=y_3=y_5$ and $y_2=k$ and $y_4=j$.
{}From these conditions it follows that move $L_u$
can be carried out $k-j$ times
to the sequence $\{(x_2,y_2),\ldots,(x_5,y_5)\}$.
This gives a new sequence
$X'=\{(x'_1,y'_1),\ldots,(x'_5,y'_5)\}$,
with $y'_1=y'_5$, $y'_2=y'_4=k$ and $y'_3=k-j$.
{}From these conditions it follows that move $L_d$
can be carried out $k-j$ times
to the sequence $\{(x'_1,y'_1),\ldots,(x'_4,y'_4)\}$.
This gives the final sequence $X''=
\{(x''_1,y''_1),\ldots,(x''_5,y''_5)\}$,
with
$y^{''}_1=y^{''}_3=y^{''}_5$, $y^{''}_2=j$ and $y^{''}_4=k$.
The total number of moves carried out is therefore $2(k-j)$.
Since in the minimal configuration
there are $n_k$ particles of charge $k$
to the left of $p_{j,1}$, this gives a total contribution
$\sum_{k=j+1}^{r-2} (k-j) \, n_k$.
Apart from this, we encounter
the situation where immediately to the left of $p_{j,1}$
we have a segment of the right-boundary particle.
In such an instant we can perform $L'_d$, moving
$p_{j,1}$ one step down. By construction of the
minimal configuration, this occurs
$\min(b-1,r-j-2)$ times.
Finally, after having descended all the way down and
having exchanged position with all particles of charge
$>j$, $p_{j,1}$ is positioned immediately to the right
of the left-boundary particle. It can then move up
exactly $\min(a-1,r-j-2)$ times using move $L'_u$.
Adding up all the contributions gives (\ref{mj}).
To see that (\ref{mj}) is unaltered by first having moved
some (or all) particles of charge greater than $j$,
consider a sequence of four points
$X=\{(x_1,y_1), (x_2,y_2), (x_3,y_3), (x_4,y_4)\}$
connected by straight lines.
First, let $y_1>y_2<y_3>y_4$ and let $p_{j,1}$ be positioned
immediately to the right of the sequence, i.e., the
origin of $p_{j,1}$ is at $(x_4,y_4)$. Also, let the
contour between the first two points not belong to the
left-boundary particle.
The total number of $L$-type steps
$p_{j,1}$ can make is then
$(y_3-y_4-j)+(y_3-y_2-j)+(y_1-y_2-j)=x_4-x_1-3j$, which is
independent of the positions of the points $(x_2,y_2)$ and
$(x_3,y_3)$. Hence carrying out any moves to $X$ does not
change the number of moves $p_{j,1}$ can make relative to $X$.
If the contour between the first two point does belong to the
left-boundary particle, this is changed to
$x_4-x_1-3j +r-2-\min(r-2-j,y_1)$ which is still independent of
the relative positions of $(x_2,y_2)$ and $(x_3,y_3)$.
Second, let $y_1<y_2>y_3<y_4$ and let $p_{j,1}$ be positioned
immediately to the left of the sequence, i.e., the
endpoint of $p_{j,1}$ is at $(x_1,y_1)$. Also, let the
contour between the last two points not belong to the
left-boundary particle.
The total number of $R$-type steps
$p_{j,1}$ can make is then
$(y_2-y_1-j)+(y_2-y_3-j)+(y_4-y_3-j)=x_4-x_1-3j$, which is
independent of the positions of the points $(x_2,y_2)$ and
$(x_3,y_3)$. Thanks to reversibility, the number of $L$-type
moves $p_{j,1}$ can make relative to $X$ is also $x_4-x_3-3j$.
If the contour between the last two point does belong to the
right-boundary particle, this again chances by a term
independent of the detailed positions of $(x_2,y_2)$ and $(x_3,y_3)$.
$\Box$
\begin{lemma}\label{lemkj}
The maximal number of $L$-type moves $p_{j,\ell}$ can make
is $k_{j,\ell-1}$, with $k_{j,\ell-1}$ the actual number
of steps taken by $p_{j,\ell-1}$
\end{lemma}
At last!, we finally encountered the fermionic nature
of our lattice-gas.
Proof:
Assume $p_{j,\ell-1}$ has made $k_{j,\ell-1}$ moves.
Obviously, (before) the first $k_{j,\ell-1}$ moves,
$p_{j,\ell}$ ``sees'' the same contour immediately
to its left as $p_{j,\ell-1}$ did, when carrying out
its leftward motion. Since
$p_{j,\ell-1}$ and $p_{j,\ell}$ are identical particles,
$p_{j,\ell}$ can thus carry out at least $k_{j,\ell-1}$ moves.
Let $p_{j,\ell}$ indeed carry out $k_{j,\ell-1}$ moves.
After that we encounter the situation of
two pure particles of charge $j$, with endpoint of
the first being origin of the next.
The right-most of the two can neither
carry out $L_u$, nor $L_d$, since
(in the notation of definition~\ref{defmoves})
$y_1=y_3$.
$\Box$
We note that the above two lemmas justify the chosen
ordering of carrying out the leftward moves.
First of all, by lemma~\ref{lemkj} it follows that
we indeed have to move $p_{j,\ell-1}$ before $p_{j,\ell}$.
Furthermore, we have to move $p_{k,m}$ before $p_{j,\ell}$,
$k>j$ since the elementary moves
only allow for leftward motion of pure particles,
see Fig.~\ref{moves}. Finally we have seen in the proof
of lemma~\ref{lemmj} that the number of moves the particles
of charge $j$ can make is independent of the actual
configuration of particles of charge $>j$.
\begin{lemma}
The contribution to the generating function $C$
of the particles of charge $j$, is given by
$\cal C$, is given by
\begin{equation}
{\cal C}_j = \Mult{m_j+n_j}{n_j}.
\label{Cj}
\end{equation}
\end{lemma}
Proof:
{}From the lemmas~\ref{genqq}, \ref{lemmj} and \ref{lemkj} we get
(dropping the subscripts $j$ in the $k$-variables)
\begin{equation}
{\cal C}_j=
\sum_{k_1=0}^{m_j} \sum_{k_2=0}^{k_1} \ldots
\sum_{k_{n_j}=0}^{k_{n_j-1}} q^{k_1+k_2+\cdots +
k_{n_j}}.
\end{equation}
We can (re)interpret this sum as the generating
function of all partitions with largest part less or
equal to $m_j$
and number of parts less or equal to $n_j$. Thus we get
(\ref{Cj}), see e.g., ref.~\cite{Andrews}.
Combining the above lemma with lemma~\ref{lemmin},
we can state our second proposition as
\begin{proposition}
The partition function of the one-dimensional Fermi-gas is
given by
\begin{equation}
Z(\vec{n};a,b)=
q^{E_{\min}} \prod_{j=1}^{r-3} {\cal C}_j=
q^{E_{\min}} \prod_{j=1}^{r-3}
\Mult{m_j + n_j}{n_j},
\label{coll}
\end{equation}
with $E_{\min}$ given by (\ref{Emin})
and $m_j$ by (\ref{mj}).
\end{proposition}
To recast this result into a simpler from, we
eliminate the $n$-variables in favour of the $m$-variables.
To do so we use the simple formulae
\begin{eqnarray}
\lefteqn{
-\min(p,q-1)+2 \,\min(p,q)-\min(p,q+1)=\delta_{p,q}
\qquad \: p,q-1 \geq 0 } \nonumber \\
\lefteqn{
-\min(p,q-1)+2 \, \min(p,q)=p+\delta_{p,q}
\qquad \qquad \quad \qquad 0\leq p \leq q+1 }
\end{eqnarray}
to get
\begin{equation}
-m_{j-1}+2m_j-m_{j+1}=
L \, \delta_{j,1} + \delta_{j,r-a-1} +
\delta_{j,r-b-1} -2 n_j
\qquad j=1,\ldots,r-3
\label{mn}
\end{equation}
with $m_0=m_{r-2}=0$. To obtain the $j=1$ case
of the above equation we made use of the completeness
relation (\ref{completeness}).
Introducing the $(r-3)$-dimensional vectors $\vec{m}$ and
$\vec{\mbox{e}}_j$ with entries $(\vec{m})_j=m_j$
and $(\vec{\mbox{e}}_j)_k=\delta_{j,k}$,
we can rewrite (\ref{mn}) as
\begin{equation}
\vec{n} = \frac{1}{2}\: (L \, \vec{\mbox{e}}_1 + \vec{\mbox{e}}_{r-a-1} +
\vec{\mbox{e}}_{r-b-1} - C \: \vec{m}).
\label{mnsystem}
\end{equation}
Substituting this into equations~(\ref{Emin}) and (\ref{coll}),
we arrive at the following simple result:
\begin{proposition}
The partition function of the Fermi-gas of content $\vec{n}$
reads
\rm
\begin{equation}
Z(\vec{n};a,b) =
q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} \,m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + L \, \vec{\mbox{e}}_1 +
\vec{\mbox{e}}_{r-a-1} +\vec{\mbox{e}}_{r-a-1})_j}{m_j},
\label{Z}
\end{equation}
\em
whth, $\vec{m}$ obtained through equation~(\ref{mnsystem}).
\end{proposition}
\subsection{Computation of $\Xi_L(a,b)$.}
Having computed the partition function of our
Fermi-gas, it is only a trivial
step to obtain the grand-canonical partition function
$\Xi_L(a,b)$, defined in (\ref{ZCab}). In particular
\begin{equation}
\Xi_L(a,b) = \sum_{\vec{n}} Z(\vec{n};a,b) .
\end{equation}
Since our final result (\ref{Z}) for $Z$ is entirely expressed
through the $m$-variables, it is natural to also
express the above sum over $\vec{n}$ in terms of a sum over
$\vec{m}$.
{}From $(\ref{mj})$, and the fact that the occupation numbers $n_j$
cannot be negative, we get
\begin{equation}
m_j = \min(a-1,r-j-2) + \min(b-1,r-j-2) + 2 \:\mbox{\sf Z} \hspace{-0.82em} \mbox{\sf Z}\,_{\geq 0}
\qquad j=1,\ldots,r-3.
\label{rm}
\end{equation}
Hence we obtain the grand-canonical partition function as
\begin{equation}
\Xi_L(a,b) =
\left. \sum_{\vec{m}}
\right.^{(0)}
q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} \,m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + L \, \vec{\mbox{e}}_1 +
\vec{\mbox{e}}_{r-a-1}
+\vec{\mbox{e}}_{r-a-1}
)_j}{m_j},
\label{Xi}
\end{equation}
where the $(0)$ in the sum over $\vec{m}$ denotes the restriction
(\ref{rm}).
\subsection{Computation of $X_L(a,b)$.}
To finally obtain the one-dimensional configuration sum
$X_L(a,b)$, we have to carry out the sum (\ref{summu}),
where we recall that $\Xi_L(a,b):=\Xi_L(a,b,0,r)$.
To get the expression for $\Xi_L(a-\mu,b-\mu,0,r-\mu)$,
we have to make the substitutions
$a\to a-\mu$, $b\to b-\mu$ and $r\to r-\mu$ in (\ref{Xi}).
This exactly gives back (\ref{Xi}) apart from the fact that
the restriction on the sum changes to
\begin{equation}
m_j = \left\{
\begin{array}{ll}
\min(a-1,r-j-2) & \\
\quad + \min(b-1,r-j-2)-2\mu + 2 \:\mbox{\sf Z} \hspace{-0.82em} \mbox{\sf Z}\,_{\geq 0} \qquad &
j=1,\ldots,r-\mu-3 \\
0 & j=r-\mu-2,\ldots,r-3.
\end{array} \right.
\label{rmmu}
\end{equation}
Denoting this restriction as $(\mu)$, we can write
\begin{equation}
X_L(a,b)=\sum_{\mu=0}^{\min(a,b)-1}
\left. \sum_{\vec{m}}
\right.^{(\mu)}
q^{\,\case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} \,m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + L \, \vec{\mbox{e}}_1 +
\vec{\mbox{e}}_{r-a-1} +\vec{\mbox{e}}_{r-a-1})_j}{m_j}.
\label{ds}
\end{equation}
Combining the sum over $\vec{m}$ restricted to $(\mu)$ and the
sum over $\mu$, gives
\begin{equation}
X_L(a,b)=
\left. \sum_{\vec{m}}
\right.^{'}
q^{\, \case{1}{4} \, \vec{m}^T C \, \vec{m}
-\case{1}{2} \,m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + L \, \vec{\mbox{e}}_1 +
\vec{\mbox{e}}_{r-a-1} +\vec{\mbox{e}}_{r-a-1})_j}{m_j},
\label{ss}
\end{equation}
with the prime denoting yet another restriction,
\begin{equation}
m_j \equiv \min(a-1,r-j-2) + \min(b-1,r-j-2)
\qquad j=1,\ldots,r-3.
\label{rmprime}
\end{equation}
Unfortunately, we have not found an elegant way to
prove this simplification and we defer it till the
appendix.
To rewrite the above form of the restriction,
in the form conjectured
in refs.~\cite{KKMMa,Melzer},
we note the identity
\begin{equation}
\begin{array}{rcccccccc}
\min(p,q) & \equiv & \delta_{p+1,q}&+&\delta_{p+3,q}&+&
\delta_{p+5,q} & + & \ldots \\
&+ & \delta_{1,q}&+&\delta_{3,q}&+&\delta_{5,q}& + &\ldots \; ,
\end{array}
\end{equation}
for $p,q\geq 0$.
Using this twice, once setting
setting $p=a-1$ and $q=r-j-2$, and once setting
$p=b-1$ and $q=r-j-2$, we get
$m_j \equiv (\vec{Q}_{a,b})_j$, with $\vec{Q}_{a,b}$ given by
(\ref{Qrest}).
We can thus
conclude this section formulating our main result as a theorem.
\begin{theorem}\label{pageth1}
\renewcommand{\arraystretch}{1}
For all $a=1,\ldots,r-1$, $b=1,\ldots,r-2$ and
$L -|a-b| \in 2 \, \:\mbox{\sf Z} \hspace{-0.82em} \mbox{\sf Z}\,_{\geq 0}$,
the one-dimensional configuration sum
(\ref{confsums}), is given by
\rm
\begin{equation}
X_L(a,b) =
\sum_{\vec{m} \equiv \vec{Q}_{a,b} }
q^{\, \case{1}{4} \,
\vec{m}^T C \, \vec{m}
-\case{1}{2} \,m_{r-a-1} }
\prod_{j=1}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + L \, \vec{\mbox{e}}_1 +
\vec{\mbox{e}}_{r-a-1} + \vec{\mbox{e}}_{r-b-1})_j}{m_j},
\nonumber
\label{Xferm}
\end{equation}
\em
where $\vec{Q}_{a,b}$ is given by (\ref{Qrest}).
\end{theorem}
\nsection{Bosonic solution of the ABF model}\label{sec4}
In this section we recall the method for computing
the sum (\ref{confsums}) to obtain (up to a prefactor)
the right-hand side of Melzer's identities (\ref{Mid}).
This alternative approach to the sum (\ref{confsums}) is
the one originally taken by Andrews, Baxter and Forrester~\cite{ABF}
and is given here mainly for reasons of completeness.
As a first step we introduce a function $Y_L(a,b)$
defined exactly as $X_L(a,b)$ in (\ref{confsums}), but
with $\sigma_{L+1}=b-2$ instead of $\sigma_{L+1}=b$.
We can then immediately infer the
recurrence relations
\begin{eqnarray}
X_L(a,b) &=& \, Y_{L-1}(a,b+1) + q^{L/2} X_{L-1}(a,b-1)
\quad \qquad 1 \leq b \leq r-2 \label{rr1} \\
Y_L(a,b) &=& X_{L-1}(a,b-1) + q^{L/2} \, Y_{L-1}(a,b+1)
\quad \qquad 2 \leq b \leq r-1,
\label{rr2}
\end{eqnarray}
subject to the initial and boundary conditions
\begin{eqnarray}
\lefteqn{
X_0(a,b) = Y_0(a,b)=\delta_{a,b}} \label{init}\\
\lefteqn{X_L(a,0)=Y_L(a,r)=0.}
\label{bound}
\end{eqnarray}
To state the solution to these equations, we quote
the following theorem established by Andrews, Baxter and
Forrester~\cite{ABF}:
\begin{theorem}
For $L\geq 0$, $1 \leq a,b,c \leq r-1$, $c=b\pm 1$, $L+a-b \equiv 0$,
let $X_L(a,b,c) := X_L(a,b)$ if $c=b+1$ and
$X_L(a,b,c) := Y_L(a,b)$ if $c=b-1$. Then
\begin{eqnarray}
\lefteqn{
X_L(a,b,c) = q^{(a-b)(a-c)/4} \sum_{j=-\infty}^{\infty} \left\{
q^{j\bigl(r(r-1)j+r(b+c-1)/2-(r-1)a\bigr)}
\Mults{L}{\frac{1}{2}(L+a-b)-rj}\right. }
\nonumber \\
& & \left. \qquad \qquad \qquad \qquad \qquad \qquad \;
-q^{ \bigl((r-1)j+(b+c-1)/2\bigr)\bigl(rj+a\bigr)}
\Mults{L}{\frac{1}{2}(L-a-b)-rj}\right\}.
\label{the2}
\end{eqnarray}
\end{theorem}
To proof this, we note
that (\ref{the2}) satisfies (\ref{rr1}), thanks to
\begin{equation}
\Mult{N}{m}
= \Mult{N-1}{m-1}+q^{m}\Mult{N-1}{m}.
\label{qexp1}
\end{equation}
Similarly,
the proof that (\ref{the2}) satisfies (\ref{rr2}) follows by
application of
\begin{equation}
\Mult{N}{m}=\Mult{N-1}{m}+q^{N-m}\Mult{N-1}{m-1} .
\label{qexp2}
\end{equation}
To show that the initial condition (\ref{init}) holds,
note (\ref{qpoly}) as well as the range of $a$ and $b$.
This gives $j=0$ as the only non-vanishing
term in the sum, and hence $X_0(a,b,c) = \delta_{a,b}$.
Finally, $X_L(a,0)=0$ follows from (\ref{the2}) upon
substitution of $b=0$, $c=1$ and making the change
of variables $j\to -j$ in the first term
withing the curly braces. Analogously, $Y_L(a,r)$ follows from
(\ref{the2}) upon substituting $b=r$, $c=r-1$
and making the change of variables $j\to -j-1$ in the first term within
the braces. $\Box$.
To obtain the desired expression for $X_L(a,b)$, we set
$c=b+1$ in (\ref{the2}) yielding
\begin{eqnarray}
\lefteqn{
X_L(a,b) = f_{a,b}^{-1} \sum_{j=-\infty}^{\infty} \left\{
q^{j\bigl(r(r-1)j+rb-(r-1)a\bigr)}
\Mults{L}{\frac{1}{2}(L+a-b)-rj} \right. }
\nonumber \\
& & \left. \qquad \qquad \qquad \qquad \;
-q^{ \bigl((r-1)j+b\bigr) \bigl(rj+a\bigr)}
\Mults{L}{\frac{1}{2}(L-a-b)-rj} \right\}.
\label{cor2}
\end{eqnarray}
Combined with theorem~\ref{pageth1} on page~\pageref{pageth1},
this proves
Melzer's polynomial identities (\ref{Mid}).
In conclusion to this section we make two remarks about the
solution (\ref{cor2}).
First, the introduction of the auxiliary function $Y_L(a,b)$
could have been avoided, since from the definition (\ref{confsums})
one can obtain recurrences that only
involve the function $X_L(a,b)$. In particular,
\begin{equation}
X_L(a,b) =\left\{
\begin{array}{ll}
q^{L/2} X_{L-1}(a,b-1) + q^{(L-1)/2} X_{L-1}(a,b+1) & \\
\qquad \qquad \qquad \qquad \, + (1-q^{L-1}) X_{L-2}(a,b) \quad &
b=1,\ldots,r-3 \label{rr3} \\
q^{L/2} X_{L-1}(a,b-1) + X_{L-2}(a,b) & b=r-2,
\end{array} \right.
\end{equation}
with the same conditions on $X_L(a,b)$ as in (\ref{init})
and (\ref{bound}).
The price to be paid for this is that, in order
to show (\ref{cor2})
solves these relations, we need double application of (\ref{qexp1})
and (\ref{qexp2}). Interestingly though, in terms of the fermionic
left-hand side of (\ref{Mid}), the recurrences (\ref{rr3})
seem to be more natural, see e.g., ref.~\cite{Berkovich}.
A second remark we wish to make is that like the fermionic
result (\ref{Mid}), also (\ref{cor2}) has a nice interpretation
in terms of restricted lattice paths.
To see this, note that in order to obtain the generating
function for the restricted lattice paths,
we can first compute the generating function
$G_L(\emptyset)$
of restricted lattice paths without the restriction
$0\leq y\leq r-2$.
Since all paths that
go below $y=0$ and above $y=r-2$ have now incorrectly been
included, we have to
subtract the generating function $G_L(\downarrow)$ of paths
that somewhere go below $y=0$, as well as the generating function
$G_L(\uparrow)$ of paths that somewhere go above $y=r-2$.
However, we are again in error, since paths that go below $y=0$
as well as above $y=r-2$ have been subtracted twice.
To correct this we add $G_L(\downarrow,\uparrow)$
and $G_L(\uparrow,\downarrow)$, being the generating function
of all paths that somewhere go above $y=r-2$ {\em after} having
gone below $y=0$ and the generating function of all paths
that somewhere go below $y=0$ {\em after} having gone
above $y=r-2$. Again this is no good, and we keep continuing
the process of adding and subtracting generating functions.
In formula this reads
\begin{equation}
X_L(a,b) = \sum_{j\geq 0} \Bigl\{
G_L(\underbrace{\downarrow\uparrow\cdots\downarrow\uparrow}_{2j})
+
G_L(\underbrace{\uparrow\downarrow\cdots\uparrow\downarrow}_{2j+2})
-
G_L(\underbrace{\downarrow\uparrow\cdots\uparrow\downarrow}_{2j+1})
-
G_L(\underbrace{\uparrow\downarrow\cdots\downarrow\uparrow}_{2j+1})
\Bigr\},
\label{sumG}
\end{equation}
with
$G_L(\underbrace{\downarrow\uparrow\cdots\downarrow\uparrow}_{2j})$
the generating function of all lattice paths that contain a sequence
of extrema
$\{(x_1,y_1),(x_2,y_2),\ldots,(x_{2j},y_{2j})\}$, with
$x_j>x_k$ for $j>k$,
$y_{2k-1}<0$ and $y_{2k}>r-2$, and with the other generating functions
defined similarly.
Of course, since we consider paths of fixed, finite length,
the above series only contains a finite number of nonzero terms.
Computing the functions $G_L$, we obtain
\begin{eqnarray}
G_L(\underbrace{\downarrow\uparrow\cdots\downarrow\uparrow}_{2j})
&=& f_{a,b}^{-1} \: q^{j\bigl(r(r-1)j-rb+(r-1)a\bigr)}
\Mults{L}{\frac{1}{2}(L-a+b)-rj} \nonumber \\
G_L(\underbrace{\uparrow\downarrow\cdots\uparrow\downarrow}_{2j})
&=& f_{a,b}^{-1} \: q^{j\bigl(r(r-1)j+rb-(r-1)a\bigr)}
\Mults{L}{\frac{1}{2}(L+a-b)-rj} \nonumber \\
G_L(\underbrace{\downarrow\uparrow\cdots\uparrow\downarrow}_{2j+1})
&=& f_{a,b}^{-1} \: q^{ \bigl((r-1)j+b\bigr) \bigl(rj+a\bigr)}
\Mults{L}{\frac{1}{2}(L-a-b)-rj} \label{compG} \\
G_L(\underbrace{\uparrow\downarrow\cdots\downarrow\uparrow}_{2j+1})
&=& f_{a,b}^{-1} \: q^{ \bigl((r-1)(j+1)-b\bigr) \bigl(r(j+1)+a\bigr)}
\Mults{L}{\frac{1}{2}(L+a+b)-r(j+1)} ,
\end{eqnarray}
for all $j\geq 0$.
Substitution in (\ref{sumG}) correctly reproduces the expression
(\ref{cor2}).
We remark that the above described method for computing
$X_L(a,b)$ is merely a rewording of the {\em sieving} technique
developed by Andrews in the context of partition theory,
see e.g., ref.~\cite{Andrews}. For the details of the calculation
leading to (\ref{compG}) we refer the reader to ref.~\cite{Andrews},
Chapter~9, and ref.~\cite{ABBBFV}.
\nsection{Summary and discussion}
In this paper we have, using the combinatorial
technique developed in part I, computed all
one-dimensional configuration sums of the $(r-1)$-state
ABF model. In contrast to the earlier results of
Andrews, Baxter and Forrester, our expressions are
of so-called fermionic type, and provide a new
proof of polynomial identities conjectured by Melzer.
In the limit of an infinitely large lattice, these
identities imply the fermionic
expressions for the $\chi_{b,a}^{(r-1,r)}(q)$
Virasoro characters as conjectured by the
Stony Brook group.
Using the Andrews--Bailey construction, we also
proved fermionic expressions
for several non-unitary minimal Virasoro characters.
In conclusion to this paper we make a few comments.
First, motivated by the ground breaking papers of the Stony Brook
group \cite{KKMMa,KKMMb,DKKMM,KM,DKMM},
a vast literature has arisen containing numerous
claims for identities of the
Rogers--Ramanujan type~[10,27-34]. We expect that
our fermionic method for computing generating
functions of restricted lattice paths can be applied to obtain
proof of several of these conjectures.
Other recently developed approaches towards
either proof, or an increase of understanding, of
Fermi-Bose character identities,
can for e.g., be found in refs.~[6,7,11,35-47].
A second remark is that in the $q=1$ limit, Melzer's
identities reduce to identities for the
number of $L$-step paths on the A$_{r-1}$ Dynkin
diagram, with fixed initial and final position.
Viewed this way, it turns out that Melzer's identities
are in fact a special 1-dimensional case of polynomial
identities for $q$-deformed path-counting on arbitrary
$d$-dimensional cuboids. In the limit of infinitely
long paths, these ``cuboid'' identities decouple
into products of Virasoro character identities.
The simplest example beyond Melzer's case is the
$q$-deformation of a path-counting formula on
a ``railroad'' of length $r-3$, reading:
\begin{eqnarray}
\lefteqn{
f_{a,b} \sum_{\vec{m} \equiv \vec{Q}_{a,b}}
q^{\: \case{1}{4} \, \vec{m}^T C \: \vec{m}
-\case{1}{2} \, m_{r-a-1} }
\Mult{L}{m_1}
\prod_{j=2}^{r-3}
\Mult{\frac{1}{2}({\cal I} \, \vec{m} + \vec{\mbox{e}}_{r-a-1}
+\vec{\mbox{e}}_{r-b-1})_j}{m_j} } \nonumber \\
& & =
\sum_{j=-\infty}^{\infty} \left\{
q^{j\big(r(r-1)j+rb-(r-1)a\big)}
\Mults{L}{2(r-1)j+b-a}_2
-q^{\big((r-1)j+b\big)\big(rj+a\big)}
\Mults{L}{2(r-1)j+b+a}_2
\right\},
\end{eqnarray}
for $a,b=1,\ldots,r-2$ and $L\geq 0$.
Here $\Mults{N}{m}_2$ are $q$-deformed trinomial coefficients
defined as \cite{AB}
\begin{equation}
\renewcommand{\arraystretch}{1.5}
\Mult{N}{m}_2 =
\sum_{k\geq 0} q^{k(k+m)}
\Mult{N}{k} \Mult{N-k}{k+m}.
\end{equation}
A discussion for the case of arbitrary cuboids
will be presented elsewhere~\cite{cuboid}.
A final remark is that the result (\ref{Xferm}) proven in this paper
has nice partition theoretical interpretations.
One follows from the work of Andrews {\em et al.}~\cite{ABBBFV},
stating that the one-dimensional configuration sum $X_L(a,b)$
is the generating function of all partitions into at most
$\case{1}{2}(L+a-b)$ parts, each part $\leq \case{1}{2}(L-a+b)$,
such that the hook differences on the $(1-b)$-th diagonal are
$\geq b-a+1$ and on the $(r-b-2)$-th diagonal $\leq b-a$.
Another interpretation follows by viewing a restricted lattice path
with total energy $E$
as a partition of $2 E=\lambda_1+\lambda_2+\ldots + \lambda_M$, with
$\lambda_j$ the $j$-th $x$-position counted from the right,
where the path has no extremum.
With this map from paths to partitions, $X_L(a,b;q^2)$
is the generating function of partitions into parts
$\lambda_1,\lambda_2,\ldots,\lambda_M$, with $\lambda_M<\ldots <\lambda_2<\lambda_1\leq L$,
and
\begin{equation}
\begin{array}{l}
1-b \leq u_j -d_j \leq r-b-2 \qquad \quad \forall j=1,\ldots,M \\
u_M-d_M = a-b,\; a-b-1.
\end{array}
\end{equation}
Here $u_j$ is the number of parts $\lambda_k \equiv a-b+k$ for $k\leq j$
and
$d_j$ is the number of parts $\lambda_k \not\equiv a-b+k$ for $k\leq j$.
\section*{Acknowledgements}
I wish to thank Alexander~Berkovich, Omar~Foda,
Peter~Forrester and Barry~McCoy
for many interesting and helpful discussions.
This work is supported by the Australian Research Council.
\section*{Note added}
After completing this manuscript we received a
preprint by A.~Schilling~\cite{Schilling},
in which (\ref{Mid}) as well
as (\ref{Mid}) are proven as special cases of polynomial
identities for finitized branching functions
of the cosets
|
\section{Introduction}
The discovery of the inverse scattering method \cite{Miura} was a
real breakthrough in theory of the classical completely integrable
Hamiltonian systems, which goes back to the classical papers of Euler,
Lagrange, Liouville, Jacobi and others. The systematic way to
construct and solve completely integrable Hamiltonian systems using
the theory of Lie groups and their representations originated in the
works of Kostant \cite{K}, Adler \cite{Ad} and Symes \cite{Symes}; it
was further developed by many other authors.
The invention of Lie-Poisson groups by Drinfeld \cite{DrLP} made it
possible to develop the general concepts underlying the theory of
classical
integrable systems and the integration of the corresponding equations
of motions \cite{SemenovDressing}. We refer the reader to review
papers \cite{Re-Sem}, \cite{Tenlect} and to the books \cite{FT},
\cite{Newell} most related to our discussion. One of the main general
results in the theory is the construction of integrable Hamiltonian
systems possessing a Lie (Lie-Poisson in general) group of symmetries
and the expression of their solutions in
terms of the solution to the factorization problem on this group
\cite{SemenovDressing}. The concrete classes of models differ by the
type of symmetry group and by the type of factorization.
Within this approach, the fundamental methods (inverse scattering
method, algebro-geometric methods of solution) and the fundamental
notions of the soliton theory, such as $\tau$-function \cite{JM} and
Baker-Akhieser function \cite{DKN}, \cite{SW}, found their unifying and
natural group-theoretical explanation.
The theory of integrable models of quantum mechanics and quantum field
theory also made a remarkable progress within the quantum version of
the inverse scattering method, which goes back to the seminal Bethe
ansatz for solving the Heisenberg
spin chain. We refer the reader, for a review of related topics, to the
books \cite{Gaudin}, \cite{Korepin} or to the papers \cite{F}, \cite{KS},
\cite{Thacker}. This
development made it possible to introduce quantum groups \cite{drinfeld},
\cite{J}
algebraic objects, playing in the quantum case a role analogous to that of
the Lie groups in the classical theory. However, we were still
missing (with the exception of the quantum integrable systems with
discrete time evolution \cite{Resh}) the quantum analogue of the
factorization theorem for the solution to the Heisenberg equations of
motion of a quantum integrable system. However, we have to mention the
remarkable paper \cite{Maillet} in this relation.
This paper is an attempt to formulate a quantum version of the
factorization theorem. As in the classical case, there is a simple direct
proof, and also a more conceptual proof which gives a generalization of the
classical
construction (based on symplectic reduction)
due to Semenov-Tian-Shansky \cite{SemenovDressing} and to which we devote the
main text. We hope that this construction has an interest of its own.
It could be, for example, interesting in relation with the Bethe ansatz
(Proposition 4), and it might be useful for a proper
formulation of the quantum version of the ${\tau}$-function. We prefer to
describe the direct proof in Appendix 1.
Section 2 contains the construction of a quantum dynamical system on
a dual quasi-triangular Hopf algebra
$F$, with the Hamiltonian $h$ taken as an arbitrary co-commutative
function on $F$, and gives its Lax pair formulation. Section 3
introduces a larger quantum dynamical system on the corresponding
Heisenberg double $D_H$ with a very simple time evolution, such that
the original quantum dynamical system (under some additional
assumptions) can be identified with a reduction of it. Section 4
contains our main result concerning the solution of the quantum Lax
equation of Section 2. Section 5 gives the formulation of our
results in a form suitable for integrable quantum chains or (after
performing the continuous limit) integrable quantum field theories.
Appendix 1 is devoted to a direct proof of our main theorem.
Finally in Appendix 2, we give a possible formulation of the factorization
problem in the case of factorizable Hopf algebras.
The paper is written for physicists. So, for example, we are working formally
with a notion of a dual Hopf algebra, which would need more detailed
specification in the infinite-dimensional case. Further, all algebraic tensor
products used in the paper would have to be properly completed in the
infinite-dimensional case.
Correspondingly we do
not discuss the precise sense in which the universal elements, such as
R-matrix, T-matrix,etc., exist. Apart from this, all constructions of the paper
are still valid. For these subtleties
and for more information about quantum groups we refer the reader to the
existing monographs on the subject (e.g. \cite{Pressleyatal}).
\section{Quantum Lax pairs}
The starting point of the following investigation is the
quasi-triangular Hopf algebra $U$ and its dual Hopf algebra $F = U^*$.
We shall use the standard notation: $m, \Delta$, $S$ and $\varepsilon$
for product, coproduct, antipode and co-unit, respectively, in both
$U$ and $F$, and also
the notation $\sum x_{(1)}\otimes x_{(2)}$ for the result of
coproduct $\Delta$ applied to $x$, in $U$ or $F$.
We start from
the commutation relation \cite{FRT}
\begin{equation}
R_{12} T_{1} T_{2} = T_{2} T_{1} R_{12} \,,\label{br}
\end{equation}
where $R \in U \otimes U$ is the universal R-matrix and $T$ is the
universal element in $U \otimes F$ (sometimes called universal
T-matrix).
In the following we will always use the notation like
$$T_1 T_2 \equiv T_{13} T_{23} \,,$$
so that, for instance, (\ref{br}) means an equality in $U\otimes U\otimes F$
and the indices $1$ and $2$ refer to the different copies of $U$ in
this triple tensor product.
We want to study the following quantum dynamical system on the
quantum group $F$. The Hamiltonian $h$ is taken to be a co-commutative
element in $F$; it holds
\begin{equation}
\Delta (h) = \sigma \Delta (h) \,,
\end{equation}
where $\sigma$ is the flip operation.
The set of all such elements form a commutative subalgebra in $F$
\cite{drinfeld}.
The quantum dynamics is given by the following Heisenberg equations
of motion:
\begin{eqnarray}
i \dot{T_2}& = &[ h , T_2 ] \cr
&= &\langle h\otimes id , T_1 T_2 - T_2 T_1 \rangle \cr
&=& \langle h\otimes id , T_1 T_2 - T_2 T_1 -
{(R_{12}^{\pm})}^{-1}T_1T_2 + T_2 {(R_{12}^{\pm})}^{-1}T_1 \rangle\,,
\end{eqnarray}
where we used the commutation relations on $F$ with the universal
elements
$R^+ = R_{21}$ and $R^- = {R_{12}}^{-1}$ and the co-commutativity of
$h$, and where the time derivative applies
to the second tensor-factor of $T\in U\otimes F$ belonging to $F$.
Therefore we can state the following proposition, generalizing the discussion
of \cite{Maillet}:
\proclaim Proposition 1.
The Heisenberg equations can be written in the Lax form
\begin{eqnarray}
i \dot{T} &=& [ M^{\pm} , T ] \,, \hskip 0.5cm \mbox{where} \cr
M_2^{\mp} &=& \langle h \otimes id , ((1- {(R_{12}^{\pm})}^{-1})
T_1\rangle \,.\label{Lp}
\end{eqnarray}
In order to construct a solution for this set of equations we
shall consider in the next sections a quantized version of the
construction by
Semenov-Tian-Shansky in \cite{SemenovDressing}.
\section{Dynamics on the quantum Heisenberg double}
The quantum Heisenberg double $D_H$ (quantum cotangent bundle of $F$)
is a smash product algebra $D_H=U{\triangleright\!\!\!<} F$ \cite{Sweedler} defined
with the help of the left action of $U$ on $F$.
We shall use the following description of $D_H$ \cite{Sem}, \cite{Zum}:
\proclaim Proposition 2.
The Heisenberg double $D_H$ is defined by
the following relations
\begin{eqnarray}
R_{12} T_1T_2 &=& T_2 T_1 R_{12} \,, \cr
R_{21} L_1^{\pm} L_2^{\pm} &=& L_2^{\pm} L_1^{\pm} R_{21} \,, \cr
R_{21} L_1^+ L_2^- &=& L_2^- L_1^+ R_{21} \,,\cr
L_1^{\pm} T_2 &=& T_2R_{12}^{\pm}L_1^{\pm} \,.
\end{eqnarray}
The universal T-matrix $T$ as well as universal L-matrices
$L^{\pm}=R^{\pm}$ are understood as elements of $U\otimes D_H$ in the
above equalities.
We shall introduce one more element of $U\otimes D_H$, denoted as $Y$
and defined as
$$Y=L^+(L^-)^{-1}.$$
Now let us consider the quantum dynamical system on the Heisenberg
double $D_H$
with the Hamiltonian ${\cal H}$ chosen to be a Casimir of $U \subset
D_H$ of the form \cite{GZB}
\begin{equation}
{\cal H}=\mbox{\mbox{Tr}}^v_1 ( Y_{12}^{-1} D_1)\,, \label{Ham}
\end{equation}
where $D\in U$ is defined, with the help of the universal R-matrix
$R=\sum R^{(1)}\otimes R^{(2)}$, as $D=\sum R^{(1)}S(R^{(2)})$ and the
superscript $v$ indicates the trace in the first factor of $U\otimes U
\subset U\otimes D_H$ evaluated
in an arbitrary representation $v$ of $U$.
The Heisenberg equations on $D_H$ take the form
\begin{eqnarray}
\dot Y &=& 0\,, \cr
i\dot T &=& T\xi^{\cal H}, \label{te}
\end{eqnarray}
with
\begin{equation}
\xi^{\cal H} =
\mbox{\mbox{Tr}}^v_1(R_{12}^{-1}Y_1^{-1}R_{21}^{-1}D_{1}-R_{12}^{-1}Y_
1^{-1}R_{12}D_{1})\,.
\end{equation}\label{xi}
Again the time derivative in (\ref{te}) applies to the second
tensor-factors of
$T$ and $Y$ belonging to $D_H$.
Since $\xi^{\cal H}\in U\otimes U \subset U\otimes D_H$ is
evidently time-independent, these Heisenberg equations are solved
trivially
\begin{eqnarray}
Y(t)&=& Y(0)\,, \cr
T(t) &=& T(0)\mbox{exp}(-it\xi^{\cal H})\,. \label{ts}
\end{eqnarray}
In the following we will assume that $U$ itself is a quantum double
$D(U_-)$
of some Hopf algebra $U_-$.
Therefore we have
$$
U = U_-\otimes U_+\,,\hskip 1cm \mbox{with}\,U_+ = U_-^{* op \Delta}
$$
as a linear space and coalgebra. Similarly , we can write
$$
F = F_-\otimes F_+\,, \hskip 1cm \mbox{with}\,F_{\pm} = U^*_{\pm}
$$
as a linear space and an algebra.
Correspondingly we have $R\in U_-\otimes U_+$ and the universal
element $T$ factorizes in $U \otimes F $ as \cite{FRT}
\begin{eqnarray}
T &=& \Lambda Z \,, \cr
\mbox {with}
\,\Lambda &\in& U_- \otimes F_- \,,\cr
\mbox {and}\, Z &\in& U_+ \otimes F_+ \,.\label{dec}
\end{eqnarray}
The commutation relations of the elements $Y$ and $Z$ assumed as
elements in $U\otimes D_H$ play a crucial role in the following.
They are given by the following lemma.
\proclaim Lemma 1.
The elements $Y$ and $Z$ commute in the following way
\begin{eqnarray}
R_{21}Y_1R_{12}Y_2&=&Y_2R_{21}Y_1R_{12}\,, \cr
R_{12}Z_1Z_2&=&Z_2Z_1R_{12}\,, \cr
Z_1 Y_1 Z_2 &=& Z_2 Z_1 Y_1 R_{12} \,.\label{zy}
\end{eqnarray}
{\em Proof.}
Only the last assertion is non-trivial. We shall omit the details of
the proof of this relation, which follows immediately from the
discussion of \cite{Jurco-Schlieker} (all arguments given there we need are
valid also in
the general situation of the present paper), if we keep in mind the
difference in the decomposition of the universal
T-matrix used there and the decomposition (\ref{dec}). The resulting
difference is that the element $Q$ used in
\cite{Jurco-Schlieker} does not appear in the commutation relations
at all.
In order to make contact with the quantum dynamical system described
in Section 2, we need the following proposition.
\proclaim Proposition 3.
There exists an embedding of $F\hookrightarrow D_H$, which is an
algebra homomorphism, given by
\begin{equation}
\tilde{T} = Z Y^{-1} Z^{-1} \,.\label{emb}
\end{equation}
This means that the relation
\begin{equation}
R_{12} \tilde{T_1} \tilde{T_2} = \tilde{T_2} \tilde{T_1} R_{12} \,
\end{equation}
holds in $U\otimes H_D$.
We shall use the symbol $\tilde F$ for the image of this embedding.
{\em Proof.}
The proof is straightforward using the commutation relations
(\ref{zy}).
This embedding of the original quantum group $F$ in the Heisenberg
double
will be used later on to project down a solution (\ref{ts}) to the
Heisenberg equations (\ref{te})
in $H_D$ to a solution of the Lax equation (\ref{Lp}) on the original
quantum phase space $F$. For doing this the following
identification of the
Hamiltonians of the corresponding systems is important.
Our starting Hamiltonian $h$ on the quantum group $F$ of Section
2 was supposed to be a co-commutative element in $F$. In the case when $F$
as its own left comodule decomposes to a direct sum of all its
irreducible comodules (a coarse form of the Peter-Weyl theorem)
the most general co-commutative element
$h$ is of the form
\begin{equation}
h=\mbox{\mbox{Tr}}^v T\,, \label{ham}
\end{equation}
where the trace in the first factor of $T\in U\otimes F$ is taken in
an appropriate representation $v$ of $U$. For simplicity, we shall
assume in the following our Hamiltonian $h$ to be exactly of
this type.
\proclaim Proposition 4.
For any representation $v$ of $U$ the equality
\begin{equation}
\mbox{\mbox{Tr}}^v_1 (Y_1^{-1} D_1) = \mbox{\mbox{Tr}}^v_1 ( Z_1
Y_1^{-1} Z_1^{-1}) = \mbox{\mbox{Tr}}^v_1 ( \tilde{T_1}) \,
\end{equation}
holds in $D_H$.
Roughly speaking the embedding (\ref{emb}) sends the trace of $\tilde
T$
in any representation of $U$ to the quantum trace of $Y$ in the same
representation, and we can identify the Hamiltonian $h$ with the reduction
to the $\tilde F$ of the Hamiltonian ${\cal H}$.
{\em Proof.} It holds in any representation $v$ of $U$ that
\begin{equation}
\mbox{\mbox{Tr}}_1 (\hat{R}_{12}^{-1} D_1 ) = 1 \,.\label{d-id}
\end{equation}
Here and in the rest of the proof we assume that both copies of $U$
to which indices
$1$ and $2$ refer are taken in the representation $v$.
{}From the third relation in (\ref{zy}) we get
\begin{equation}
Y_1^{-1} Z_1^{-1} Z_2 P_{12} = Z_2\hat{R}_{12}^{-1} Y_2^{-1}
Z_2^{-1}
\,,
\end{equation}
where $P_{12}$ is the permutation operator in the representation $v$.
Now taking the quantum trace of this equation and using (\ref{d-id})
we obtain
\begin{equation}
\mbox{\mbox{Tr}}^v_1 Y_1^{-1} Z_1^{-1} Z_2 P_{12} D_1 = Z_2 Y_2
Z_2^{-1} \label{*} \,.
\end{equation}
Taking now the usual trace in the second tensor-factor (and
renaming the tensor-factors) yields the desired identity
\begin{equation}
\mbox{\mbox{Tr}}^v_1 (Y_1^{-1} D_1 ) = \mbox{\mbox{Tr}}^v_1 (Z_1
Y_1^{-1} Z_1^{-1}) \,.
\label{trace}
\end{equation}
\section{Solution to the Lax equation}
Now we can return to our dynamical system, on $\tilde T\in U\otimes
D_H$ governed
by the Hamiltonian ${\cal H}$ of the form (\ref{Ham}), of Section 3. There, we
constructed the solution (\ref{ts}) to the equations
of motion
of this quantum dynamical system; we showed that
there exists an embedding of the original quantum group $F$ in the
Heisenberg double $D_H$, such that the Hamiltonians of the corresponding
systems coincide after this embedding.
In this chapter we are going to use it to obtain a
solution to the quantized Lax equations (\ref{Lp}) on the quantum group $F$.
Let us denote as $g(t)$ the following element
$$g(t)=Z(0)\mbox{exp}(-it\xi^{\cal H})Z(0)^{-1}\in U\otimes D_H.$$
Now, the time evolution on $D_H$ is an algebra homomorphism, and so the
decomposition of $T(t)$ in $U \otimes F(t)$ in the same form as in
(\ref{dec})
makes sense:
\begin{eqnarray}
T(t) &=& \Lambda(t) Z(t)\,,\cr
\mbox {with}\,
\Lambda &\in& U_- \otimes F_-(t) \,,\cr
\mbox{and}\, Z &\in& U_+ \otimes F_+(t) \,.\label{dec1}
\end{eqnarray}
So as a consequence of (\ref{ts}) the element $g(t)$ can be expressed
as
\begin{eqnarray}
g(t) &=& \Lambda(0)^{-1} T(t) Z(0)^{-1} \,,\cr
&=& \Lambda(0)^{-1} \Lambda(t)Z(t) Z(0)^{-1} \,.\label{LZ}
\end{eqnarray}
This gives us a decomposition of $g(t) \in U\otimes
D_H$, with
\begin{eqnarray}
g(t)&=&g_-(t)g_+(t), \hskip 0.5cm g_{\pm}\in U_{\pm}\otimes D_H\,,\cr
g_-(t)&=&\Lambda(0)^{-1} \Lambda(t), \hskip 0.5cm g_+(t)=Z(t)
Z(0)^{-1}. \label{decc}
\end{eqnarray}
We will now show that $g(t)$ and its factors $g_{\pm}(t)$ are actually
elements of $U\otimes \tilde F
\subset U\otimes D_H$.
Let us define
\begin{equation}
M = R_- (h_{(1)}) \otimes h_{(2)} - R_+ (h_{(1)}) \otimes h_{(2)} \in U\otimes
\tilde F,
\end{equation}
with
\begin{eqnarray}
R_+(x)&=&\sum\langle x, R^{(1)}\rangle R^{(2)}\,, \cr
R_-(x)&=&\sum S(R^{(1)})\langle x, R^{(2)}\rangle\,, \label{maps}
\end{eqnarray}
and demonstrate that
\begin {equation}
g(t)=\mbox{exp}(-itM(0)) \in U\otimes
\tilde F\,.\label{g}
\end{equation}
That this is really true follows from the definition of $\xi^{\cal
H}$ (\ref{xi}), co-commutativity of $h$
and the following chain of identities:
\begin{eqnarray}
\mbox{\mbox{Tr}}_1^v(R_{12}^{-1}Y_1^{-1}(R_{12}^{\pm})^{-1}D_1)&=&Z_2^
{-1}Z_2\mbox{\mbox{Tr}}_1^v(R_{12}^{-1}Y_1^{-1}(R_{12}^{\pm})^{-1}D_1)
\,, \cr
&=&Z_2^{-1}\mbox{\mbox{Tr}}_1^v(Y_1^{-1}Z_1^{-1}Z_2Z_1(R_{12}^{\pm})^{
-1}D_1)\,,\cr
&=&Z_2^{-1}\mbox{\mbox{Tr}}_1^v(Y_1^{-1}Z_1^{-1}(R_{12}^{\pm})^{-1}Z_1
D_1)Z_2\,,\cr
&=&Z_2^{-1}\mbox{\mbox{Tr}}_1^v(Z_1Y_1^{-1}Z_1^{-1}(R_{12}^{\pm})^{-1}
)Z_2\,,\label{le}
\end{eqnarray}
where we used successively the third and the second relations of (\ref{zy}) and
the
relation
(\ref{*}).
Let us mention that $M=M^+-M^-$, with $M^{\pm}$ the elements of
$U\otimes \tilde F$
entering the Lax equation (\ref{Lp}).
{}From the equality $\xi^{\cal H}=Z^{-1}MZ$ that we just proved, and from the
time independence of $\xi^{\cal H}$, we have
$$ M(t) = g_+(t)M(0)g_+(t)^{-1}. $$
Writing now
\begin{equation}
g_+(t)=Z(t)Z(0)^{-1}
=\mbox{exp}(-it(1\otimes{\cal H}))Z(0)\mbox{exp}(it(1\otimes{\cal H}))
Z(0)^{-1}
\end{equation}
and using the last equality (with a $-$ sign) in (\ref{le}) and Proposition 4
we get immediately
\begin{equation}
g_+(t)=\mbox{exp}(-it(1\otimes h)\mbox{exp}(-it(M^+(0) -1\otimes
h)).\label{fga}
\end{equation}
For $g_-(t)$ we get similarly
\begin{equation}
g_-(t)=\mbox{exp}(-it(1\otimes h-M^-(0)))\mbox{exp}(it(1\otimes h)),\label{fgb}
\end{equation}
which follows from (\ref{fga}) and (\ref{commzero}) (in Appendix 1).
This shows that indeed $g_{\pm}\in U_{\pm}\otimes \tilde F$, as we claimed.
Moreover $g_{\pm}$ are the unique solutions of the equations
\begin{equation}
i\dot g_+=M^+g_+ \label{M+}.
\end{equation}
and
\begin{equation}
i\dot g_-=-g_-M^-,\label{M-}
\end{equation}
with initial condition $g_{\pm}(0)=1$.
Starting now from:
\begin{eqnarray}
\tilde T (t) &=& Z(t) Y^{-1}(0) Z(t)^{-1}\,,\cr
&=& Z(t)Z(0)^{-1}\tilde T(0)Z(0) Z(t)^{-1}\,,\label{tau}
\end{eqnarray}
we arrive at the main result of this paper:
\proclaim Theorem 1. Let $U$ be a quasi-triangular Hopf
algebra and let $F$ be its dual Hopf algebra. Let $g(t)$ be given by
(\ref{g}), with the Hamiltonian $h$, taken to be any co-commutative element of
$F$, and let $U_{\pm}$ denote the ranges of the mappings $R_{\pm}$
(\ref{maps}). Then $g(t)$ can be factorized:
\begin{equation}
g(t)=g_-(t)g_+(t)\,,\label{fac}
\end{equation}
$g_{\pm}(t)\in U_{\pm}\otimes F$ given by (\ref{fga}), (\ref{fgb}). Moreover
$g_{\pm}(t)$ are the unique solutions of
equations (\ref{M+}), (\ref{M-}), with initial conditions $g_{\pm}(0)=1$.
The element $T(t) \in U\otimes F$, given by
\begin{equation}
T(t)= g_+(t)T(0)g_{+}(t)^{-1} = g_-(t)^{-1}T(0)g_-(t) \,,\label{main}
\end{equation}
solves the quantum Lax equation (\ref{Lp}). In the case of factorizable $U$ we
can interpret (\ref{fac})
as a well-formulated factorization problem in $U\otimes F$ (see Appendix 2).
Although we proved here Theorem 1 only in the special case of $U$ being
a quantum double and the Hamiltonian $h$ being of the form (\ref{ham}), we
formulated it more generally. We shall give a simple direct proof of Theorem
1 in full generality in
Appendix 1.
The second equality in (\ref{main}) is due to fact that $g(t)$
commutes with $T(0)$ in $U\otimes F$, which is easily seen, e.g. from
(\ref{Lp}).
To specify completely our quantum dynamical system, we have to choose
a representation $\pi$ of the quantum group $F$. The algebra of
quantum observables will be the image $\pi(F)$ of $F$ in the chosen
representation.
The time evolution of an observable $\pi(a)$, $a \in F$, will then be
given by
$\pi(a)(t)=\pi((\langle a\otimes id), T(t)\rangle)$.
\section{Lax equations for quantum chains}
In this section we will discuss how the above result modifies in the
case of a quantum spin chain. The algebra of observables
$F^{\otimes N}$ for the chain consists of $N$ independent copies
$F^n,\,n=1,2,...,N$, of the dual Hopf algebra $F$ of a
quasi-triangular Hopf algebra $U$. We also ste $N+1\equiv 1$.
We will denote as $L^n \in U\otimes F^n$ the copy of the
universal T-matrix corresponding to the site $n$. We reserve the
character $T$ for the quantum monodromy matrix
$$T= L^1 ...L^N\,\in U\otimes F^{\otimes N}\,.$$
Then we have the following relations in $U\otimes F^{\otimes N}$
\begin{eqnarray}
R_{12} L_1^i L_2^i &=& L_2^i L_1^i R_{12} \,,\cr
L_1^i L_2^j &=& L_2^j L_1^i \,,\hskip 0.5cm i\neq j \,. \label{commi1}
\end{eqnarray}
The quantum monodromy matrix
satisfies
\begin{equation}
R_{12} T_1 T_2 = T_2 T_1 R_{12}\,
\end{equation}
and for the partial products
$$
{\psi}^n = L^1 ...L^{n-1} \,, \hskip 0.5cm {\psi}^1=1\,,
$$
we obtain
\begin{equation}
R_{12} {\psi}_1^n {\psi}_2^n = {\psi}_2^n {\psi}_1^n
R_{12}\,.\label{commi2} \end{equation}
We will choose our Hamiltonian $h\in F^{\otimes N}$ as any element of
$F^{\otimes N}$ of the form
\begin{equation}
h=\langle (H\otimes id),T \rangle\,, \label{chaham}
\end{equation}
with co-commutative $H\in F$. Again such elements form a commutative
subalgebra in $F^{\otimes N}$.
\proclaim Proposition 5.
The Lax equations for the site $n$ have the following
form:
\begin{eqnarray}
i\dot{L}^n &=& M^{\pm n} L^n - L^n M^{\pm (n+1)}, \hskip 0.5cm
\mbox{where} \cr
M_2^{\mp n} &=& \langle H \otimes id , ((1- (R_{12}^{\pm})^{-1})
(\psi_1^n)^{-1} T_1 \psi_1^n \rangle \,. \label{cLp}
\end{eqnarray}
This can be easily shown using the commutation relations
(\ref{commi1}), (\ref{commi2}) and co-commutativity of $H$ in the
same way as in Proposition 1.
Lax pair of Proposition 5 formalizes concrete examples of Lax
pairs known
for particular integrable quantum chains or integrable field
theoretical models
\cite{Korepin1}, \cite{Sklyanin}, \cite{Kulish-Sklyanin},
\cite{Maillet}, \cite{Zhang}, \cite{Sogo}.
To avoid a cumbersome notation we introduce again a notation
similar to that in the previous section:
$$\hat M= R_- (H_{(1)}) \otimes H_{(2)} - R_+ (H_{(1)}) \otimes H_{(2)}\in
U\otimes F$$
and
$$ M=\langle \hat M\otimes id, id\otimes T\rangle \in U\otimes
F^{\otimes N}.$$
The folowing modification of Theorem 1 can be proved analogically
as in the previous sections. The twisted Heisenberg double of
ref. \cite{Sem} should be used for this.
However, there is also a direct proof using Theorem 1.
\proclaim Theorem 2.
Let $U$ be a quasi-triangular Hopf algbera, $F$ its dual Hopf algebra.
Let us assume a quantum chain system as described above with the
Hamiltonian $h$ given in (\ref{chaham}), where $H$ is any co-commutative
element
of $F$. Then the elements $g^n(t)\in U\otimes F^{\otimes N}$:
\begin{equation}
g^n(t) =({\psi}^n(0))^{-1} \mbox{exp}(-itM(0)){\psi}^n(0)\,,\label{gn}
\end{equation}
can be decomposed as
\begin{equation}
g^n(t)=g^n_-(t)g^n_+(t)\,,
\end{equation}
with $g^n_{\pm}(t)\in U_{\pm}\otimes F^{\otimes N}$ given by
\begin{eqnarray}
g^n_+(t)&=&\mbox{exp}(-it(1\otimes h)\mbox{exp}(-it(M^{+n}(0) -1\otimes
h))\,,\cr
g^n_-(t)&=&\mbox{exp}(-it(1\otimes h-M^{-n}(0)))\mbox{exp}(it(1\otimes
h)).\label{fgl}
\end{eqnarray}
Moreover $g^n_{\pm}$ are the unique solutions of equations (\ref{M+}),
(\ref{M-}) (all entries indexed by $n$), with initial conditions
$g^n_{\pm}(0)=1$. The elements
$L^n(t)\in U\otimes F^{\otimes N}$
\begin{equation}
L^n (t)= g_+^n (t) L^n (0) (g_+^{n+1} (t))^{-1} = (g_-^n (t))^{-1}
L^n (0) g_-^{n+1} (t)
\end{equation}
solve the chain Lax equations (\ref{cLp}). In the case of factorizable $U$,
elements $g_{\pm}$ can be thought of
as a solution to the factorization problem for $g$ as formulated in Appendix 2.
{\em Proof.}
Following the same reasoning as led to Proposition 1, we can establish
that the Heisenberg equations of motion for entries of the quantum monodromy
matrices
$$T^n=(\psi^n)^{-1} T \psi^n=L^n...L^NL^1...L^{n-1},$$
for chains obtained from the original one by a shift $(1,...,N)\mapsto
(n,...,N,1,...,n-1)$, are precisely of the form (\ref{Lp}), with
$M^{\pm}=M^{\pm n}$. So the time evolution of the quantum monodromy matrix
$T^n$ is given by Theorem 1, with $g(t)=\mbox {exp}(-it(M^{+n}(0)-M^{-n}(0))$,
which means that all elements $\mbox {exp}(-it(M^{+n}(0)-M^{-n}(0))\in U\otimes
F^{\otimes N}$ can be decomposed as claimed.
It remais only to show that $$\mbox
{exp}(-it(M^{+n}(0)-M^{-n}(0))=({\psi}^n(0))^{-1}
\mbox{exp}(-itM(0)){\psi}^n(0).$$
This is, however, a consequence of the co-commutativity of $H$ and the
following
equality:
\begin{equation}
\psi_1^n(R_{12}^{\pm})^{-1}
(\psi_1^n)^{-1} T_1=(\psi_2^n)^{-1}(R_{12}^{\pm})^{-1}
T_1\psi_2^n,
\end{equation}
which easily follows from (\ref{commi1}), (\ref{commi2}).
The rest is trivial.
\vskip 0.3cm
In this paper we did not mention the dressing symmetries of the quantum
integrable systems at all. However, dressing symmetries
can be introduced in a way completely analogous to the classical case (for the
classical case see \cite{SemenovDressing}). This aspect of the theory
of quantum integrable systems will be discussed elsewhere.
\vskip 1cm
\noindent{\bf Acknowledgements}
The authors would like to thank Bruno Zumino
for many valuable discussions. B.J. wishes also to acknowledge
discussions
with H. Grosse, P. Kulish and N. Reshetikhin. He would like
to thank Professors Grosse and Zumino for their kind hospitality at
ESI and LBL, respectively.
\vskip 1cm
\noindent
{\bf Appendix 1: Direct proof of Theorem 1}
\vskip 0.6cm
As in the classical case there is a simple direct proof of Theorem 1
and hence also of Theorem 2, which as it follows from the discussion of the
preceding section, is a simple consequence of the Theorem 1.
Let $U$ be any quasi-triangular Hopf algebra, $F$ its dual Hopf algebra and
$U_{\pm}$ the range of the maps $R_{\pm}$ (\ref{maps}).
First of all let us mention that $g_{\pm}\in U_{\pm}\otimes F$, given by
(\ref{fga}) and (\ref{fgb}):
\begin{eqnarray*}
g_+(t)&=&\mbox{exp}(-it(1\otimes h)\mbox{exp}(-it(M^+(0) -1\otimes h))\,,\cr
g_-(t)&=&\mbox{exp}(-it(1\otimes h-M^-(0)))\mbox{exp}(it(1\otimes h)).
\end{eqnarray*}
solve the equations (\ref{M+}), (\ref{M-}) with initial condition
$g_{\pm}(0)=1$, for any co-commutative Hamiltonian $h$. This is easily
checked by a direct computation.
As an immediate consequence we find that $T(t)$ given by (\ref{main})
solve the Lax equations (\ref{Lp}).
Now we shall show that the elements
$R_- (h_{(1)}) \otimes h_{(2)}\in U_-\otimes F$ and $R_+ (h_{(1)}) \otimes
h_{(2)}\in U_+\otimes F$ commute.
Let us compute
\begin{eqnarray*}
R_{20}^{-1}T_0R_{12}T_1&=&
R_{20}^{-1}R_{12}T_0T_1=R_{20}^{-1}R_{12}R_{10}T_1T_0R_{10}^{-1}\cr
&=&R_{10}R_{12}R_{20}^{-1}T_1T_0R_{10}^{-1}=R_{10}R_{12}T_1R_{20}^{-1}T_0R_{10}^{-1}.
\end{eqnarray*}
Dualizing the first and last term in the above chain of equalities, which take
place in $U\otimes U\otimes U\otimes F$, in the components $0$ and $1$ with
$h\otimes h\in F\otimes F$, and using the co-commutativity of the Hamiltonian
$h$,
we have
\begin{equation}
[R_- (h_{(1)}) \otimes h_{(2)},R_+ (h_{(1)}) \otimes h_{(2)}]=0.
\label{commzero}
\end{equation}
This shows that
\begin{equation}
g_-(t)g_+(t)=\mbox{exp}(-it(M^+(0)-M^-(0))).
\end{equation}
So we have proved Theorem 1 directly.
\vskip 1cm
\noindent
{\bf Appendix 2: Factorization problem}
\vskip 0.6cm
Here we make an attempt to formulate a quantum analogue of the factorization
problem from the classical case \cite{SemenovDressing} in the case where $U$ is
a factorizable Hopf algebra \cite{S-R}. Similarly to
\cite{S-R}, we can give in this case an equivalent description of
the algebra structure of the tensor product $U\otimes F$. We shall omit
details.
The claim is that as a linear space $U\otimes F=F^{(-)}\otimes F^{(+)}$, where
$F^{(\pm)}$ are subalgebras of $U\otimes F$, both as algebras isomorphic to
$F$. They are embedded into $U\otimes F$ via the
the following algebra morphisms:
\begin{eqnarray*}
{\cal R}_{\pm}: F&\hookrightarrow& U\otimes F\,, \cr
x&\mapsto&R_{\pm}(x_{(1)}) \otimes x_{(2)}\,,
\end{eqnarray*}
where $R_{\pm}$ are given by (\ref{maps}).
This vector space isomorphism can be made into an algebra isomorphism if the
commutation relations between the elements of the two copies $F^{(\pm)}$ of $F$
are introduced through
\begin{equation}
(1\otimes x)(y\otimes 1)= \langle R_{21}^{-1},y_{(1)} \otimes x_{(1)}\rangle
y_{(2)} \otimes x_{(2)}\langle R_{21},y_{(3)} \otimes x_{(3)}\rangle\,,
\end{equation}
so that, as an algebra, $U\otimes F$ is isomorphic to a bicrossproduct of two
copies of $F$.
This means that any element $\alpha\in U\otimes F$ can be expressed as
\begin{equation}
\alpha = \sum \alpha_{i-}\alpha_{i+},
\end{equation}
with all $\alpha_{i-}$ lying in the range of the map ${\cal R}_-$ and all
$\alpha_{i+}$ lying in the range of the map ${\cal R}_+$, respectively. All
${\alpha_{i\pm}}$ are given unambiguously.
It may happen that some particular $\alpha\in U\otimes F$, if expressed
in this way, is a simple product of two factors
\begin{equation}
\alpha = \alpha_-\alpha_+,
\end{equation}
$\alpha_+$ being the image under ${\cal R}_+$ of a (unique)
invertible element $x\in F$ and $\alpha_-$ being image under ${\cal R}_-$ of
the inverse $x^{-1}$ of the same element $x$.
If this is the case, we shall refer to the unique elements
\begin{eqnarray}
\tilde \alpha_-&=&\alpha_-(1\otimes x)\,,\cr
\tilde \alpha_+&=&(1\otimes x^{-1})\alpha_+
\end{eqnarray}
as to the solution of the factorization problem for $\alpha \in U\otimes F$.
Clearly the elements $g_{\pm}$ and $g_{\pm}^n$ from Theorems 1 and 2 are, in
the case
of the factorizable $U$, solutions to the factorization problem for $g$ and
$g^n$, respectively.
Finally we have to note that in concrete examples it is possible to give an
alternative characterization of the factorization of elements $g$ or $g^n(t)$.
We shall
discuss this very briefly for typical example when our starting Hopf algebra is
the quantum double of a Yangian $Y$ \cite{drinfeld}: $U=D(Y)$. Other cases
are similar. Let $T_{\lambda}$ be the automorphism of $Y$ of \cite{drinfeld}.
We shall use the same notation $T_{\lambda}, \lambda\in$ {\sr C} for its
extension (via duality) to the full double.
Then the decomposition of $(T_{\lambda}\otimes id)g^n(t)$,
$$(T_{\lambda}\otimes id)g^n(t)=(T_{\lambda}\otimes
id)g_-^n(t)(T_{\lambda}\otimes id)g_+^n(t),$$
is uniquely determined by the assumption that $(T_{\lambda}\otimes
id)g_{\pm}^n(t)$ are regular as functions of $\lambda$ in {\sr C}$P_1\backslash
\{\infty\}$
and {\sr C}$P_1\backslash\{0\}$, respectively, and $(T_{\infty}\otimes
id)g_+^n=1$.
|
\section{#1}\renewcommand{\theequation}
{\mbox{\arabic{section}.\arabic{equation}}}\setcounter{equation}{0}}
\renewcommand{\ss}[1]{\subsection{#1}}
\newcommand{\sss}[1]{\subsubsection{#1}}
\newcommand{\app}[1]{\section{#1}\renewcommand{\theequation}
{\mbox{\Alph{section}.\arabic{equation}}}\setcounter{equation}{0}}
\renewcommand{\date}[1]{\par\bigskip\par\sl\hfill #1\par\medskip\par\rm}
\newcommand{\email}[1]{e-mail: \sl #1@science.unitn.it\rm}
\newcommand{\femail}[1]{\footnote{\email{#1}}}
\newcommand{\pacs}[1]{\smallskip\noindent{\sl PACS number(s):
\hspace{0.3cm}#1}\par\bigskip\rm}
\def\hrule\par\begin{description}\item{Abstract: }\it{\hrule\par\begin{description}\item{Abstract: }\it}
\def\par\end{description}\hrule\par\medskip\rm{\par\end{description}\hrule\par\medskip\rm}
\newcommand{\ack}[1]{\par\section*{Acknowledgments} #1}
\renewcommand{\vec}[1]{{\bf #1}}
\def\M{{\cal M}}
\newcommand{\ca}[1]{{\cal #1}}
\def\hs{\qquad\qquad}
\def\nn{\nonumber}
\def\beq{\begin{eqnarray}}
\def\eeq{\end{eqnarray}}
\def\ap{\left.}
\def\at{\left(}
\def\aq{\left[}
\def\ag{\left\{}
\def\cp{\right.}
\def\ct{\right)}
\def\cq{\right]}
\def\right\}} %%% close {\right\}}
\newtheorem{Theorem}{Theorem}
\newtheorem{Lemma}{Lemma}
\def\R{\mbox{$I\!\!R$}}
\def\N{\mbox{$I\!\!N$}}
\def\Z{\mbox{$Z\!\!\!Z$}}
\def\C{\mbox{$I\!\!\!\!C$}}
\def\ii{\infty}
\def\X{\times\,}
\newcommand{\fr}[2]{\mbox{$\frac{#1}{#2}$}}
\def\tr{\,\mbox{tr}\,}
\def\Tr{\,\mbox{Tr}\,}
\def\PP{\,\mbox{PP}\,}
\def\Res{\,\mbox{Res}\,}
\def\ach{\,\mbox{cosh$^{-1}$}\,}
\def\ash{\,\mbox{sinh$^{-1}$}\,}
\def\ath{\,\mbox{tanh$^{-1}$}\,}
\def\acth{\,\mbox{coth$^{-1}$}\,}
\renewcommand{\Re}{\,\mbox{Re}\,}
\renewcommand{\Im}{\,\mbox{Im}\,}
\def\lap{\Delta}
\def\cc{\phi_c}
\def\alpha{\alpha}
\def\beta{\beta}
\def\gamma{\gamma}
\def\delta{\delta}
\def\varepsilon{\varepsilon}
\def\zeta{\zeta}
\def\iota{\iota}
\def\kappa{\kappa}
\def\lambda{\lambda}
\def\varrho{\varrho}
\def\sigma{\sigma}
\def\omega{\omega}
\def\varphi{\varphi}
\def\theta{\theta}
\def\vartheta{\vartheta}
\def\upsilon{\upsilon}
\def\Gamma{\Gamma}
\def\Delta{\Delta}
\def\Lambda{\Lambda}
\def\Sigma{\Sigma}
\def\Omega{\Omega}
\def\Theta{\Theta}
\def\Theta{\Theta}
\def\Upsilon{\Upsilon}
\begin{document}
\preprint{UTF 357}
\title{
Finite Temperature Effects for Massive Fields \\
in $D$-dimensional Rindler-like Spaces}
\author{Andrei A. Bytsenko\footnote{email: <EMAIL>
(subject: Prof. A.A. Bytsenko)}}
\address{State Technical University, St. Petersburg 195251, Russia}
\author{Guido Cognola\femail{cognola} and
Sergio Zerbini\femail{zerbini}}
\address{\dinfn}
\date{August 1995}
\hrule\par\begin{description}\item{Abstract: }\it
The first quantum corrections to the free energy for massive
fields in $D$-dimensional space-times of the form
$\R\times\R^+\times\M^{N-1}$, where $D=N+1$ and $\M^{N-1}$
is a constant curvature manifold, is investigated by means of the
$\zeta$-function regularization.
It is suggested that the nature of the divergences, which are
present in the thermodynamical quantities, might be better
understood making use of the conformal related
optical metric and associated techniques. The general form of the
horizon divercences of the free energy is obtained as a function of
free energy densities of fields having negative square masses (absence
of the gap in the Laplace operator spectrum) on ultrastatic manifolds
with hyperbolic spatial section $H^{N-2n}$ and of the Seeley-DeWitt
coefficients of the Laplace operator on the manifold $\M^{N-1}$.
Furthermore, recurrence relations are found
relating higher and lower dimensions.
The cases of Rindler space,
where $\M^{N-1}=\R^{N-1}$ and very massive $D$-dimensional black holes,
where $\M^{N-1}=S^{N-1}$ are treated as examples.
The renormalization of the internal energy is also discussed.
\par\end{description}\hrule\par\medskip\rm
\pacs{04.62.+v, 04.70.Dy}
\s{Introduction}
Recently there has been a renewed interest in the physics of black
holes. Several issues like the interpretation of the
Bekenstein-Hawking classical formula for the black hole entropy, the puzzle
of loss of information in the black hole evaporation
and the interpretation of the Hawking temperature have been discussed
(see, for example, the review \cite{beke95b}).
Furthermore in many papers, it has been pointed
out that it should be desirable
to have the usual statistical interpretation of the
black hole entropy as the number of the gravitational states at the
horizon and to try to understand the dynamical origin of the black
hole entropy (see, for example \cite{barv95-51-1741}).
On general grounds, the density of levels as a function of
the mass $M$, of a $D$-dimensional black hole should read
\cite{harm93-47-2438} (for details, see Sec. 6.3)
\beq
\Omega(M)\simeq C_D(M)
\exp\at\frac{4\pi\hat{G}_D}{D-2}M^{\frac{D-2}{D-3}}\ct
\:.\label{td}
\eeq
where $C_D(M)$ is a quantum prefactor and $\hat G_D$ is related
to the generalized Newton constant (see Eq.~(\ref{rh})).
For a 4-dimensional black hole we have \cite{thoo85-256-727}
\beq
\Omega(M)\simeq C_4(M)\exp(4\pi G M^2)
\:,\label{t}
\eeq
$4\pi GM^2$ being the Bekenstein-Hawking classical entropy
\cite{beke73-7-2333,hawk75-43-199,gibb77-15-2752}.
As 't Hooft has pointed out, the prefactor $C_4(M)$ (the first
quantum correction to the classical result), which is usually computable for
quantum fields or extended objects (such as strings or p-branes)
in an ultrastatic space-time background, turns out to be divergent.
This prefactor may be regarded as the contribution
associated with the first quantum correction to the classical free energy.
Several different methods have been used in dealing
with such divergences, for example "the brick wall method"
\cite{thoo85-256-727,suss94-50-2700,ghos94-73-2521}, the conical
singularity method
\cite{bana94-72-957,dowk94-11-55,solo94u-246,furs95-10-649,call94-333-55,kaba94-329-46},
critically discussed in \cite{empa95-51-5716,deal95u-33}
and the related "entaglement entropy method"
\cite{bomb86-34-373,sred93-71-666,lars94u-89}.
The horizon divergences have also been associated with the information loss
issue of black holes \cite{thoo85-256-727,suss94-50-2700} and their
physical origin, for quantum fields
\cite{isra76-57-107,barb94-50-2712} or strings
\cite{dabh95-347-222,eliz95-10-1187,empa94u-3,byts95u-130},
may be described by the following simple considerations.
In a $D$-dimensional static space-time with horizons,
the equivalence principle implies that a system in thermal
equilibrium has a local Tolman temperature given by
$T(x)=T/{\sqrt {-g_{00}(x)}}$, $T$ being the asymptotic temperature.
Roughly speaking, near the canonical horizon
(this means that the quantity $g_{00}$ has simple zeros),
a static space-time may be regarded as a Rindler-like space-time.
We will show that, if one denotes by $\rho$ the proper
distance from the horizon, one gets for the Tolman temperature
$T(\rho)=T/\rho$. As a consequence, the total entropy for a quantum
bosonic gas reads (omitting a multiplicative constant)
\beq
S\equiv\int d{\vec x}\int_0^\ii T(\rho)^{D-1}\,d\rho \simeq
A T^{D-1}\int_0^\ii\rho^{-D+1}\,d\rho
\:,\label{mmm}
\eeq
where $A$ is the integral on the transverse coordinates $\vec x$,
namely the area of the horizon.
The latter integral is clearly divergent.
Introducing a horizon cutoff parameter $\varepsilon$ we may rewrite it as
\beq
S&\simeq & AT\int_\varepsilon^\Lambda \rho^{-1}\,d\rho
\simeq AT\ln \frac{\Lambda}{\varepsilon} \, , \hs\mbox{for }D=2\,\, \Lambda\,\,
\mbox{infrared cutoff} \nn \\
S&\simeq & AT^{D-1}\int_\varepsilon^\ii\rho^{-D+1}\,d\rho\simeq
\frac{AT^{D-1}}{(D-2)\varepsilon^{D-2}}\, , \hs\mbox{for } D>2
\:.\label{nnn}\eeq
For the sake of generality, we write down the asymptotic high
temperature expansion for the entropy of a quantum gas
on a $D=N+1$-dimensional static space-time defined by the metric
\beq
ds^2=g_{00}(\vec{x})(dx^0)^2+g_{ij}(\vec{x})dx^idx^j\:,
\hs \vec{x}=\{x^j\}\:,\hs i,j=1,...,N\:,
\label{x}
\eeq
$g(\vec{x})$ denoting its determinant.
Again the equivalence principle leads to
\beq
S\simeq
T^{N}\int\at\frac{g(\vec{x})}{g_{00}(\vec{x})}
\ct^{-N/2}\,dx^{N}\,.
\label{k}
\eeq
As a consequence, the horizon divergences depend on the
nature of the poles of the integrand $(g/g_{00})^{-N/2}$.
In general, for extremal black holes, where $g_{00}$ has higher order
zeroes, the divergences are much more severe than in the non extremal
case (see for example Refs.~\cite{mitr95u-42,deal95u-33,cogn95u-348}).
These considerations suggest the use of the optical metric
$\bar{g}_{\mu\nu}=g_{\mu\nu}/g_{00}$, conformally
related to the original one, in order to
investigate these issues. It is our
opinion that the conformal transformation techniques are particularly
suitable for studying finite temperature effects for fields in
space-times with horizons and here we would like to present some examples
of computation. This method has already been appeared for example in
Refs.~\cite{dowk78-11-895,page82-25-1499,brow85-31-2514} and has been
recently used in the horizon divergence problems in
Refs.~\cite{barb94-50-2712,barv95-51-1741,empa95-51-5716,deal95u-33,barb95u-155,bord94u-54}.
See also \cite{deal94u-347}, where the same
result is obtained with a different approach.
One of the purposes of this paper is to implement this idea in the case of
massive fields in $D$-dimensional Rindler-like space-times, we are going
to introduce.
Let us consider static space-times admitting canonical horizons and
having the topology of the form $\R\times\R^+\times\M^{N-1}$.
The metric reads
\beq
ds^2=-\frac{b^2\rho^2}{r_H^2} dx_0^2+d\rho^2+d\sigma^2_{N-1}\, ,
\label{rl}
\eeq
where $r_H$ is a dimensional constant, $b$ a constant factor and
$d\sigma^2_{N-1}$ the spatial metric related to the
$N-1$-dimensional manifold $\M^{N-1}$.
If $\M^{N-1}\equiv\R^{N-1}$, $b=1$ and $r_H=1/a$,
$a$ being a constant acceleration, one has to deal with the Rindler space-time.
Quantum fields in such a case have been considererd in many places,
see for example
Refs.~\cite{cand76-17-2101,haag84-94-219,free85-255-693,birr82b,more95u-52}.
If $\M^{N-1}\equiv S^{N-1}$, $b=(D-3)/2$ and $r_H$ is the Schwarzschild
radius of a black hole, then we shall show that one is dealing
with a space-time which approximates, near the horizon and in the
large mass limit, a $D$-dimensional black hole (see Sec.~\ref{S:BH}).
It is well known that space-times with
canonical horizons admit a distinguished temperature, the
(Unruh) Hawking temperature. There are several ways to compute it, one
of the simplest makes use of the relation with the related surface
gravity. The other one consists in imposing the absence of conical
singularities in the Euclidean continuation of the space-time
itself \cite{gibb77-15-2752}.
For the metric (\ref{rl}), one obtains
\beq
\beta_H=\frac{2 \pi r_H}{b}\,.
\label{ht}
\eeq
One can arrive at the same result working without using the Euclidean
continuation method, but making use of the principle of local
definiteness in quantum field theory \cite{haag84-94-219,more95u-50}.
Note that these method are no longer equivalent when one is dealing
with extremal black holes \cite{ghos95u-32}.
It is also important to stress that the variable $\rho$ defined by the
metric in Eq.~(\ref{rl}) has the meaning of radial proper distance
between the horizon and a point outside it and so, the divergences
of thermodynamical quantities are automatically expressed in an
invariant way.
The contents of the paper are the following.
In Sec.~\ref{S:CTT}, a review of the
necessary conformal transformation techniques is presented.
In Sec.~\ref{S:SF}, we consider a Laplace-type
operator defined on a class of $D=N+1$-dimensional space-times,
whose spatial sections have metrics
conformally related to $\M^N=\R^+\times\M^{N-1}$.
Since in general $\M^N$ in a non compact manifold,
the Laplace operator has a continuum spectrum and a general form
for the Plancherel measure, which is the analogue of the
degeneracy in the case of discrete spectrum, is presented.
The measure is used in Sec.~\ref{S:HK} in order to obtain a useful
form for the trace of the heat kernel, which is necessary for the
derivation of the free energy, which we derive in Sec.~\ref{S:FE}.
It is pointed out that in the Rindler case,
the spatial section of the conformally related space-time
turns out to be an $N$-dimensional hyperbolic manifold. In this case,
the massless scalar field can be treated without approximations.
In Sec.~\ref{S:PA} some applications
to the statistical mechanics of a scalar field
in $D$-dimensional Rindler and black hole space-times are presented
and the divergences of the first quantum corrections
to free energy and entropy are given.
Finally, we end with some conclusions in Sec.~\ref{S:C}
and with a resume of heat kernel, $\zeta$-function and
free energy on constant curvature manifolds in the Appendix.
\s{Conformal transformation techniques and optical manifold}
\label{S:CTT}
In this section we shall briefly summarize the method of
conformal transformations using $\zeta$-function regularization
\cite{dowk78-11-895,gusy87-46-1097,dowk88-38-3327,dowk89-327-267}.
These techniques permit to compute all physical quantities in an ultrastatic
manifold (called the optical manifold \cite{gibb78-358-467}) and,
at the end of calculations, transform back them to a static one.
This method is particularly useful in
dealing with finite temperature effects for quantum fields, since these
effects can be easily investigated in ultrastatic space-times.
To start with, we consider a non self-interacting scalar field
on a $D=N+1$-dimensional static space-time defined by the metric
(\ref{k}), i.e.
\beq
ds^2=g_{00}(\vec{x})(dx^0)^2+g_{ij}(\vec{x})dx^idx^j\:,
\hs \vec{x}=\{x^j\}\:,\hs i,j=1,...,N\:.
\eeq
The one-loop partition function is given by (we perform the Wick
rotation $x_0=-i\tau$, thus all the differential operator one is
dealing with will be elliptic)
\begin{equation}
Z=\int d[\phi]\,
\exp\at-\frac12\int\phi L_D \phi d^Dx\ct
\:,\end{equation}
where $\phi$ is a scalar density of wight $-1/2$ and
the operator $L_D$ has the form
\beq
L_D=-\lap_D^g+m^2+\xi R^g
\:.\eeq
Here $m$ (the mass) and $\xi$ are arbitrary parameters, while
$\lap_D^g$ and $R^g$ are respectively the Laplace-Beltrami operator
and the scalar curvature of the manifold in the original metric $g$.
The ultrastatic metric $\bar{g}_{\mu\nu}$ can be related to the static one by
the conformal transformation
\beq
\bar{g}_{\mu\nu}(\vec{x})=e^{2\sigma(\vec{x})}g_{\mu\nu}(\vec{x})
\:,\eeq
with
$\sigma(\vec{x})=-\frac{1}{2}\ln g_{00}$. In this manner,
$\bar{g}_{00}=1$ and $\bar{g}_{ij}=g_{ij}/g_{00}$ (optical metric).
Recalling that by a conformal transformation (we remind that $\phi$ is
a scalar density)
\beq
R^{\bar g}&=&e^{-2\sigma}\aq R^g-2(D-1)\lap_D^g\sigma
-(D-1)(D-2)g^{\mu\nu}
\partial_{\mu}\sigma\partial_{\nu}\sigma\cq\:,\nn\\
\bar\phi&=&e^{\sigma}\phi\:,\\
\lap_D^{\bar g}\bar\phi&=&e^{-\sigma}\aq\lap_D^g
-\frac{D-2}{2}\lap_D^g\sigma
-\frac{(D-2)^2}{4}g^{\mu\nu}
\partial_{\mu}\sigma\partial_{\nu}\sigma
\cq\phi\nn\\
&=&e^{-\sigma}\aq\lap_D^g+\xi_D(e^{2\sigma}R^{\bar g}-R^g)
\cq\phi\nn
\:,\label{CT}\eeq
one obtains
\beq
L_D\phi=e^{\sigma}\ag-\lap_D^{\bar g}+\xi_DR^{\bar g}
+e^{-2\sigma}[m^2+(\xi-\xi_D)R^g]\right\}} %%% close \bar\phi
\:,\eeq
where $\xi_D=(D-2)/4(D-1)$ is the conformal invariant factor.
{}From the latter equation we have
$\phi L_D\phi=\bar\phi\bar L_D\bar\phi$, where, by definition
\beq
\bar L_D=e^{-\sigma}L_De^{-\sigma}=-\lap_D^{\bar g}+\xi_DR^{\bar g}
+e^{-2\sigma}\aq m^2+(\xi-\xi_D)R^g\cq
\label{Aq}
\:.\eeq
This means that action $\bar I=I$ by definition.
Note that classical conformal invariance requires the action
to be invariant in form, that is $\bar I=I[\bar\phi,\bar g]$,
as to say $\bar L_D=L_D$. As is well known,
this happens only for conformally coupled massless fields ($\xi=\xi_D$).
For the one-loop partition function we have
\beq
\bar Z=J[g,\bar g]\,Z
\:,\eeq
where $J[g,\bar g]$ is the Jacobian of the conformal transformation.
Such a Jacobian can be computed for any infinitesimal conformal
transformation \cite{gusy87-46-1097}. To this aim it is
convenient to introduce a family of continuous conformal transformations
\beq
g^q_{\mu\nu}=e^{2q\sigma}g_{\mu\nu}
=e^{2(q-1)\sigma}\bar g_{\mu\nu}\:,\hs
\sqrt{g^q}\equiv\sqrt{|\det g^q_{\mu\nu}|}=e^{Dq\sigma}\sqrt{g}
\:,\eeq
in such a way that the metric is $g_{\mu\nu}$ or $\bar{g}_{\mu\nu}$
according to whether $q=0$ or $q=1$ respectively.
Then one gets
\beq
\ln J[g_q,g_{q+\delta q}]
=\ln\frac{Z_{q+\delta q}}{Z_q}
=\frac{\delta q}{(4\pi)^{D/2}}
\int k_D(x|L_D^q)\sigma(x)\sqrt{g^q}d^Dx
\label{deJ}
\:,\eeq
where $k_D(x|L_D^q)$, is the Seeley-DeWitt coefficient, which in the case of
conformal invariant theories, is proportional to the trace anomaly.
In general, one has the asymptotic expansion
\beq
\Tr e^{-tL_D}\simeq\sum_n K_n(L_D)t^{\frac{n-D}{2}}\:,\hs
K_n(L_D)=\frac1{(4\pi)^{D/2}}\int_{\M} k_n(x|L_D)\sqrt{g}\,d^Dx
\:.\label{hk}
\eeq
If the manifold is without boundary then $K_n=0$ for any odd $n$.
The heat kernel coefficients are computable and depend on invariant quantities
built up with curvature (field strength) and their derivatives (see,
for example, \cite{bran90-15-245}).
The Jacobian for a finite transformation can be obtained from
Eq.~(\ref{deJ}) by an elementary integration in $q$
\cite{gusy87-46-1097}. In particular we have
\beq
\ln J[g,\bar{g}]=\frac{1}{(4\pi)^{D/2}}
\int_0^1dq\int k_D(x|L_D^q)\sigma(x)\sqrt{g^q}\,d^Dx
\:,\label{lnJ}\eeq
and finally, making use of the $\zeta$-function
regularization, one has
\beq
\ln Z=\ln\bar Z-\ln J[g,\bar g]
=\frac{1}{2}\zeta'(0|\bar L_D\ell^2)-\ln J[g,\bar{g}]
\:,\label{lnZ-Zbar}\eeq
where $\ell$ is an arbitrary parameter necessary to adjust the
dimensions and $\zeta'$ represents the derivative
with respect to $s$ of the $\zeta$-function
$\zeta(s|\bar L_D\ell^2)$ related to the operator
$\bar L_D$, which is given by Eq.~(\ref{Aq}).
The same analysis can be easily extended to the finite temperature
case \cite{dowk88-38-3327}. In fact we recall that for a scalar field in
thermal equilibrium at
finite temperature $T=1/\beta$ in an ultrastatic space-time, the corresponding
partition function $\bar{Z}_\beta$
may be obtained, within the path integral approach, simply by Wick rotation
$\tau=ix^0$ and imposing a $\beta$ periodicity in $\tau$ for
the field $\bar\phi(\tau,x^i)$ ($i=1,...,N$, $N=D-1$)
\cite{bern74-9-3312,dola74-9-3320,wein74-9-3357,kapu89b}.
In this way, in the one loop approximation, one has
\begin{equation}
\bar{Z}_\beta=\int_{\bar\phi(\tau,x^i)=\bar\phi(\tau+\beta,x^i)}
d[\bar\phi]\,\exp\at-\int_0^\beta
d\tau\int\bar\phi\bar L_D\bar\phi\,d^Nx\ct
\:.\label{PF}
\end{equation}
in which
\beq
\bar L_D=-\partial_\tau^2-\bar\lap_N+\xi_DR^{\bar g}
+e^{-2\sigma}\aq m^2+(\xi-\xi_D)R^g\cq
=-\partial_\tau^2+\bar{L}_N
\:.\label{aconf}\eeq
Since the space-time is ultrastatic, by means of the $\zeta$-function
regularization again one easily obtain \cite{byts94u-325}
\beq
\ln\bar Z_\beta&=&-\frac{\beta}{2}\aq
\PP\zeta(-\fr12|\bar{L}_N)
+(2-2\ln2\ell)\Res\zeta(-\fr12|\bar{L}_N)\cq \nn \\
&&\hs+\lim_{s \to 0}\frac{d}{ds} \frac{\beta}{\sqrt{4\pi}\Gamma(s)}
\sum_{n=1}^\ii\int_0^\ii t^{s-3/2}\,e^{-n^2\beta^2/4t}\,
\Tr e^{-t\bar{L}_N}\,dt
\label{logPF-Jacobi}\:.
\eeq
where $\PP$ and $\Res$ stand for the principal part and the residue
of the function.
The free energy is related to the canonical partition function
by means of equation
\beq
F(\beta)=-\frac{1}{\beta}\ln Z_\beta
=-\frac{1}{\beta}\at\ln\bar Z_\beta-\ln J[g,\bar g]\ct
=F_0+F_\beta\:,\label{FE}\eeq
where $F_0$ represents the vacuum energy,
which is given by the first term in
Eq.~(\ref{logPF-Jacobi}), while $F_\beta$ represents the
temperature dependent part (statistical sum). The entropy and the
internal energy of the
system are given by the usual thermodynamical formulae
\beq
S_\beta=\beta^2\partial_\beta F(\beta)=\beta^2\partial_\beta F_\beta
\:.\label{e}\eeq
\beq
U_\beta=\beta\partial_\beta F(\beta)+ F_\beta=-e\partial_\beta \ln Z_\beta
\:.\label{ie}\eeq
In a similar way, one can consider spinor fields. It is sufficient to
make use of the following formal identity \cite{dowk89-327-267}
\beq
F_f(\beta)\equiv 2F_{2\beta}-F_\beta\:,
\label{FFE}\eeq
where on the r.h.s. the spinor quantities are left understood in the
formal expression of the bosonic free energy. As a consequence, the
horizon divergences of the bosonic sector cannot be compensated by the
corresponding fermionic ones.
\s{Spectral function for rank 1 Riemannian spaces conformally related to
$\M^N=\R^+\times\M^{N-1}$ }
\label{S:SF}
In many interesting physical cases, the Euclidean
optical metric my be written in the form (see Eq. (\ref{rl}) and Sec. 6)
\beq
d\bar{s}^2=d\tau^2+ \frac{r_H^2}{\rho^2}
\at d\rho^2+d\sigma^2_{N-1} \ct
\:,\label{rlo}\eeq
where $d\tau=b dx^0$, $r_H$ being a characteristic length (for example the
horizon radius),
$(r_H/\rho)^2$ the conformal factor and $d\sigma^2_{N-1}$ the metric of a
$N-1$-dimensional manifold .
Here we derive the spectral measure of the operator $\bar L_N$,
as defined by Eq.~(\ref{aconf}), acting on scalars in the spatial
section of the manifold defined by the metric Eq.~(\ref{rlo}).
Using such equations
(for convenience now we put $r_H=1$; in this way
all quantities are dimensionless; the dimensions will be easily
restored at the end of calculations) we easily obtain
\beq
dV&=&\rho^{-N}\,d\rho\,dV_{N-1}\:,\nn\\
\bar L_N&=&-\lap_N^{\bar g}-\varrho_N^2+C\rho^2\:,\label{da}\\
\lap_N^{\bar g}&=&\rho^2\partial_\rho^2-(N-2)\rho\,\partial_\rho
+\rho^2\lap_{N-1}\nn\:,\eeq
where $\lap_{N-1}$ is the Laplace-Beltrami operator on the
manifold $\M^{N-1}$ and $dV_{N-1}$ its invariant measure.
We have also set $\varrho_N=(N-1)/2$ and $C=m^2+\xi R^g$.
It should be noted the appearance of an effective "tachionic" mass
$-\varrho_N^2$, which has important consequences on the structure of the
$\zeta$-function related to the operator $\bar L_N$.
In order to study the quantum properties of matter fields defined on
this ultrastatic manifold, it is sufficient to investigate the kernel
of the operator $e^{-t\bar{L}_N} $.
To this aim, we will search for the spectral resolution
of the elliptic operator $\bar{L}_N$.
Let $\Psi_{r\alpha}(x)$ be its eingenfunctions, namely
\beq
\bar{L}_N \Psi_{r\alpha}(x)=\lambda_r^2 \Psi_{r\alpha}(x)
=\lambda_r^2f_{\alpha}(\vec x)\phi_{r\alpha}(\rho)
\:,\label{LNPsi}
\eeq
where $f_{\alpha}(\vec x)$ are the (normalized) eingenfunctions
of the reduced operator $L_{N-1}=-\lap_{N-1}+C$ with eigenvalues
$\lambda_\alpha^2$. Note that we assume $C$ to be constant. This means that
we restrict ourselves to consider only manifolds $\M^D$ with constant
scalar curvature, or alternatively minimally coupled fields.
Moreover, to avoid null eigenvalues we suppose $C>0$, but the
results can be easily extended to the case $C=0$.
Note that the spectrum of $L_{N-1}$ could also be continuum.
The differential equation which determines the continuum spectrum
turns out to be
\beq
\aq\rho^2\,\partial_\rho^2
-(N-2)\rho\,\partial_\rho
-\rho^2\lambda_\alpha^2+\varrho_N^2+\lambda_r^2
\cq\phi_{r\alpha}(\rho)=0
\:.\label{Bessel}\eeq
The only solutions of the latter equation
with the correct decay properties at
infinity are the Bessel functions of imaginary argument
$K_{ir}(\rho\lambda_\alpha)$ with $\lambda_r^2=r^2$
(if $C=0$ the operator $L_{N-1}$ has a zero mode and gives other
solutions to Eq.~(\ref{Bessel})).
Thus we have
\beq
\phi_{r\alpha}(\rho)=\rho^{\frac{N-1}{2}}
K_{ir}(\rho\lambda_\alpha)
\:.\label{10}\eeq
If we interpret in the sense of the distribution the
following innner product
\beq
(\Psi_{r\alpha},\Psi_{r'\alpha'})=
\int_0^\ii\frac{d\rho}{\rho^N}\int dV_{N-1}\,
\Psi^*_{r\alpha}(x)\Psi_{r'\alpha'}(x)
\:,\label{vbvb}
\eeq
we have the normalization condition
\beq
(\Psi_{r\alpha},\Psi_{r'\alpha'})=\delta_{\alpha\al'}
\frac{\delta(r-r')}{\mu(r)}
\:,\label{mu}
\eeq
where $\mu(r)$ is the spectral measure
associated with the continuum spectrum.
Thus, for the heat kernel of any suitable function $f(\bar L_N)$
we may write
\beq
<x|f(\bar L_N)|x'>
=\int_0^\ii dr\,\mu(r)\,f(r^2)
\sum_\alpha\Psi^*_{r\alpha}(x')\Psi_{r\alpha}(x)
\:.\label{hk2}
\eeq
The measure $\mu(r)$ may be determined
in the following standard way (Harish-Chandra's method \cite{helg84b}),
which makes use of the asymptotic behaviour of the Mac Donald
functions at the origin.
{}From Eq.~(\ref{Bessel}) and its complex conjugate
and making use of Eq. (\ref{10}) one arrives at
\beq
(\lambda_r^2-\lambda_{r'}^2)\at\phi_{r\alpha},\phi_{r'\alpha'}\ct
=\lim_{\rho\to 0}\rho^{-(N-2)}
\at\partial_\rho\phi^*_{r\alpha}\phi_{r'\alpha'}
-\phi^*_{r\alpha}\partial_\rho\phi_{r'\alpha'}\ct
\:.\label{v}
\eeq
By means of Eqs. (\ref{10}) and (\ref{mu}) we get
\beq
\frac{\delta(r-r')}{\mu(r)}
=\lim_{u\to0}\frac{u}{r^2-r'^2}
\aq\partial_u\,K_{ir}^*(u)K_{ir'}(u)
-K_{ir}^*(u)\partial_uK_{ir'}(u)\cq
\:,\label{v1}
\eeq
where again, the limit has to be understood in the sense of
distributions. Recalling that for $u\to0$
\beq
K_{ir}(u)\sim\frac12\aq
\Gamma(-ir)\at\frac u2\ct^{ir}
+\Gamma(ir)\at\frac u2\ct^{-ir}\cq\:,
\hs ir\not\in\Z\eeq
and
\beq
\lim_{u\to 0}\frac{u^{\pm ix}}{x}=\mp\pi i\delta(x)
\:,\label{asdf}\eeq
one finally has
\beq
\mu(r)=\frac{2}{\pi|\Gamma(ir)|^2}=\frac2{\pi^2}\,r\sinh\pi r
\:,\label{v2}
\eeq
which is in agreement with the 2-dimensional Kontorovich-Lebedeev inversion
formula \cite{terr85b}.
Since our aim is to evaluate the trace of functions of $\bar L_N$
using Eq.~(\ref{hk}), in particular $\Tr\exp(-t\bar L_N)$,
it is convenient to make the sum over $\alpha$, introducing
the total spectral measure
\beq
\mu_{\bar{L}_N}(r,x)=\mu(r)\,\rho^{N-1}\sum_\alpha|f_{\alpha}(\vec x)
K_{ir}(\rho\lambda_\alpha)|^2
\label{PM}
\eeq
and integrate on the manifold defining
\beq
\mu_I(r)=\int_{{\cal M}^N}\mu_{\bar{L}_N}(r,x)\,dV=
\mu(r)\,\int_0^\ii
\sum_\alpha |K_{ir}(\rho\lambda_\alpha)|^2\frac{d\rho}{\rho}
\:,\label{PMI}\eeq
where we have used the normalization properties of $f_\alpha$.
In this way, for any suitable function $f(\bar L_N)$ we have
\beq
\Tr f(\bar L_N)=\int_0^\ii f(r^2)\,\mu_I(r)\,dr
\:.\label{TrfLN}\eeq
As a simple application of Eq.~(\ref{PM}) let us consider a massless
scalar field in a $D$-dimensional Rindler space-time.
In this case the optical spatial section turns out to be the
hyperbolic space $H^{N}$ and the measure $\mu_{\bar L_N}$
should not depend on $x$, since one is dealing with
a homogeneous space and it should coincide with the known Plancherel measure.
For this case $\M^{N-1}=\R^{N-1}$ and moreover $C=0$.
The reduced operator $L_{N-1}=-\lap_{N-1}$ has a continuum spectrum,
the eigenvalues being $k^2=\vec k\cdot\vec k$ and the corresponding
eigenfunctions $f_{\vec k}=(2\pi)^{-(N-1)/2}\exp(i\vec k\cdot\vec x)$.
As a consequence
\beq
\Phi_N(r)\equiv\mu_{\bar L_N}(r)&=&\mu(r)\,
\frac{\Omega_{N-2}\rho^{N-1}}{(2\pi)^{N-1}}
\int_0^\ii|K_{ir}(\rho k)|^2\,k^{N-2}\,dk
\nn\\&=&
\frac{2}{(4\pi)^{N/2}\Gamma(N/2)}
\frac{|\Gamma(ir+\varrho_N)|^2}{|\Gamma(ir)|^2}
\:,\label{Plan}
\eeq
$\Omega_N$ being the volume of the $N$-dimensional sphere.
Of course, Eq.~(\ref{Plan}) is the correct Plancherel measure
of the Laplace operator in $H^N$ \cite{camp90-196-1}.
\s{The heat kernel for massive fields}
\label{S:HK}
Here we derive a general expression for $\mu_I(r)$ by making use of
Eq.~(\ref{PMI}) and then derive the trace of the heat kernel,
which is needed for the construction of the partition function according
to Eq.~(\ref{logPF-Jacobi}).
Now we use the Mellin-Barnes
representation \cite{grad80b}
\beq
K_{ir}^2(\rho\lambda_\alpha)=\frac1{4i\sqrt\pi}\int_{\Re z>1}
\frac{\Gamma(z+ir)\Gamma(z-ir)\Gamma(z)}{\Gamma(z+1/2)}
\,\rho^{-2z}\lambda_\alpha^{-2z}\,dz
\label{ml}
\eeq
and observe that, for $\Re z>(N-1)/2$, the sum over $\alpha$ can be done
and gives
\beq
\sum_\alpha\lambda_\alpha^{-2z}=\zeta(z|L_{N-1})
\:.\label{zred}\eeq
In integrating over $\rho$, one has to pay attention
to the fact that the result is formally divergent.
For this aim we introduce a horizon cutoff parameter $\varepsilon$ and,
when possible, we take the limit $\varepsilon\to0$.
Then we get
\beq
\mu_I(r)=\frac{\mu(r)}{8i\sqrt\pi}
\int_{\Re z=c>(N-1)/2}
\frac{\Gamma(z+ir)\Gamma(z-ir)\Gamma(z)\zeta(z|L_{N-1})}
{z\Gamma(z+1/2)\varepsilon^{2z}}\,dz
\:.\label{muIr}\eeq
The integration over $z$ can be done since the meromorphic structure
of $\Gamma$-and $\zeta$-functions are known. In fact, we have
\cite{mina49-1-242}
\beq
\Gamma(z)\zeta(z|L_{N-1})
=\sum_{n=0}^\ii\frac{K_n(L_{N-1})}{z-\frac{N-1-n}2}
+J_{N-1}(z)
\:,\label{S}\eeq
where $J_{N-1}$ is an analytic function.
Since the manifold $\M^{N-1}$ has no boundary,
all $K_n$ with odd $n$ are vanishing.
To make the integral we consider the rectangular contour
$\Gamma\equiv\{\Re z=c,\Im z=a,\Re z=-c,\Im z=-a\}$ and observe that the
two horizontal paths $\Im z=\pm a$ give a vanishing contribution
in the limit $a\to\ii$, as well as the path $\Re z=-c$
in the limit $\varepsilon\to0$. Also the poles in the strip $-c<\Re z<0$
give a vanishing contribution as soon as $\varepsilon\to0$.
Then we have to take into consideration only the
poles of the integrand in Eq.~(\ref{muIr})
in the half-plane $\Re z\geq0$. Such a function
has simple poles at the points $z=0$, $z=-n\pm ir$ and
$z=(N-1-n)/2$ ($n\geq0$).
If $D$ is even, that is $N$ is odd, $z=0$ is a double pole.
It is clear that all poles with $\Re z>0$
give rise to divergences, the number of them depending on $N$,
while the poles at $z=0$ and $z=\pm ir$ give rise to finite
contributions. As a result one obtains
\beq
\mu_I(r)&=&\sum_{n=0}^{\aq\frac{N-2}2\cq}
\frac{K_{2n}(L_{N-1})\,\Phi_{N-2n}(r)}{N-1-2n}
\at\frac{4\pi}{\varepsilon^2}\ct^{\frac{N-1-2n}2}
\nn\\&&\hs
+\frac{\zeta'(0|L_{N-1})}{2\pi}
+\frac{\zeta(0|L_{N-1})}{2\pi}\aq
\psi(ir)+\psi(-ir)-2\ln\frac\ep2-\pi\delta(r)\cq
\:,\label{muIrFF}\eeq
where $\aq\frac{N-2}2\cq$ is the integer part of the number $\frac{N-2}2$,
$\psi(z)$ the logarithmic derivative of $\Gamma$
and $\delta(r)$ the usual Dirac $\delta$-function.
Note that for even $N$, $\zeta(0|L_{N-1})$ is vanishing and so the last
term in the latter equation disappears.
Now the trace of the heat kernel can be computed by using
Eq.~(\ref{TrfLN}) with $f(r^2)=\exp(-tr^2)$.
We write it in the form
\beq
\Tr e^{-t\bar L_N}&=&
\sum_{n=0}^{\frac{N-3}2}
\frac{K_{2n}(L_{N-1})\,K(t|-\lap_{H^{N-2n}}-\varrho^2_{N-2n})}
{N-1-2n}\at\frac{4\pi}{\varepsilon^2}\ct^{\frac{N-1-2n}2}
\nn\\&&\hs\hs\hs
+\frac{1}{4\sqrt{\pi t}}
\aq\zeta'(0|L_{N-1})-2\zeta(0|L_{N-1})\ln\frac\ep2\cq
\nn\\&&\hs\hs
-\frac{\zeta(0|L_{N-1})}4
+\frac{\zeta(0|L_{N-1})}{2\pi}
\int_{-\ii}^{\ii}\psi(ir)e^{-tr^2}\,dr
\:,\label{Ktodd}\eeq
\beq
\Tr e^{-t\bar L_N}&=&\sum_{n=0}^{\frac{N-2}2}
\frac{K_{2n}(L_{N-1})\,K(t|-\lap_{H^{N-2n}}-\varrho^2_{N-2n})}
{N-1-2n}\at\frac{4\pi}{\varepsilon^2}\ct^{\frac{N-1-2n}2}
\nn\\&&\hs\hs\hs\hs
+\frac{1}{4\sqrt{\pi t}}\zeta'(0|L_{N-1})
\:,\label{Kteven}\eeq
valid for odd and even $N$ respectively.
Here by $K(t|-\lap_{H^{N-2n}}-\varrho^2_{N-2n})$ we indicate the
diagonal heat kernel of a Laplace-like operator on $H^{N-2n}$.
Of course, it does not depend on the coordinates since
hyperbolic manifolds are homogeneous. Such a kernel is known in any
dimension \cite{byts94u-325} (see the Appendix).
As in the previous section, as a simple application of Eq.~(\ref{muIrFF}),
we again consider a massless scalar field in the $D$-dimensional Rindler
space-time. We have
$K_0(L_{N-1})=(4\pi)^{-\frac{N-1}2}V_{N-1}$, $K_n=0$ for $n>0$ and
$\zeta(z|L_{N-1})=0$ for $z<(N-1)/2$.
Here $V_{N-1}$ is the volume of the manifold $\M^{N-1}$
(infinite transverse area). Then, using Eq. (\ref{Plan}), we immediately obtain
\beq
\mu_I(r)=\Phi_N(r)\,V_\varepsilon\:,\hs
V_\varepsilon=\frac{V_{N-1}\varepsilon^{-(N-1)}}{N-1}
\:,\eeq
which is the integral version of Eq.~(\ref{Plan}).
Here $V_\varepsilon$ may be considered as the volume of $H^N$.
\s{The thermodynamical quantities}
\label{S:FE}
Now it is quite straightforward to obtain the partition function and then
all the others thermodynamical quantities by means of
Eqs.~(\ref{logPF-Jacobi}) and (\ref{FE}).
Since the vacuum energy has been extensively studied in many papers,
here we concentrate our attention on the temperature dependent part of
the free energy (statistical sum) $F_\beta=\bar F_\beta=-\ln\bar Z_\beta /\beta$.
Using Eqs. (\ref{Ktodd}) and (\ref{Kteven}) we get
\beq
F_\beta^{even\:D}&=&
\sum_{n=0}^{\frac{N-3}2}
\frac{K_{2n}(L_{N-1})\,{\cal F}_{N-2n}^\beta}
{N-1-2n}\at\frac{4\pi}{\varepsilon^2}\ct^{\frac{N-1-2n}2}
\nn\\&&\hs
-\frac{\pi}{12\beta^2}
\aq\zeta'(0|L_{N-1})-2\zeta(0|L_{N-1})\ln\frac\ep2\cq
-\frac{\zeta(0|L_{N-1})}4\:\frac{\ln\beta}\beta
\nn\\&&\hs\hs
+\frac{\zeta(0|L_{N-1})}{2\pi\beta}
\int_0^{\ii}[\psi(ir)+\psi(-ir)][1-e^{-\beta r}]\,dr
\:,\label{FEeven}\eeq
\beq
F_\beta^{odd\:D}&=&
\sum_{n=0}^{\frac{N-2}2}
\frac{K_{2n}(L_{N-1})\,{\cal F}_{N-2n}^\beta}
{N-1-2n}\at\frac{4\pi}{\varepsilon^2}\ct^{\frac{N-1-2n}2}
-\frac{\pi}{12\beta^2}\zeta'(0|L_{N-1})
\:,\label{FEodd}\eeq
where ${\cal F}_{N-2n}^\beta$ indicates the free energy density for a
scalar field with (negative) square mass $-\varrho_{N-2n}^2$
on an ultrastatic manifold with hyperbolic $H^{N-2n}$ spatial section,
which has been studied in detail in Ref.~\cite{byts94u-325}
and is given in the Appendix.
Some remarks on Eqs.~(\ref{FEeven}) and (\ref{FEodd}) are in order.
First of all, it has to be noted that the parameter
$\beta$ is the inverse of the physical temperature
only if $b=1$. More generally, before to interpret
$\beta^{-1}$ as the temperature in Eqs.~(\ref{FEeven}) and (\ref{FEodd}),
one has to make the substitution $\beta\to b\beta$. The reason is due to
the fact that, in order to write the metric (\ref{rl}) in the form (\ref{rlo}),
we have changed the time coordinate according to $\tau=b x_0$.
Independently on the manifold $\M^{N-1}$, we see that the (non renormalized)
free energy has a leading divergence of the kind $\varepsilon^{-(D-2)}$
proportional to the transverse area $V_{D-2}$,
since $K_0$ and ${\cal F}_N^\beta$ are always non vanishing.
More generally, one has $\aq\frac{D-1}2\cq$ divergences of the kind
$\varepsilon^{-(D-2-2n)}$ (depending on the manifold and the operator
$L_{N-1}$) and, for even $D$, also a possible logarithmic
divergence. All these divergences are also present in the expressions
of internal energy and entropy and their expressions can be obtained by
means of Eqs. (\ref{ie}) and (\ref{e}). For example the internal
energy reads
\beq
U_\beta^{even\:D}&=&
\sum_{n=0}^{\frac{N-3}2}
\frac{K_{2n}(L_{N-1})\,{\cal U}_{N-2n}^\beta}
{N-1-2n}\at\frac{4\pi}{\varepsilon^2}\ct^{\frac{N-1-2n}2}
\nn\\&&\hs
+\frac{\pi}{12\beta^2}
\aq\zeta'(0|L_{N-1})-2\zeta(0|L_{N-1})\ln\frac\ep2\cq
+\frac{\zeta(0|L_{N-1})}{4\beta}
\nn\\&&\hs
-\frac{\zeta(0|L_{N-1})}{2\pi}
\int_0^{\ii} r[\psi(ir)+\psi(-ir)]e^{-\beta r}\,dr+U_0(\varepsilon)
\:,\label{Eeven}\eeq
\beq
U_\beta^{odd\:D}&=&
\sum_{n=0}^{\frac{N-2}2}
\frac{K_{2n}(L_{N-1})\,{\cal U}_{N-2n}^\beta}
{N-1-2n}\at\frac{4\pi}{\varepsilon^2}\ct^{\frac{N-1-2n}2}
+\frac{\pi}{12\beta^2}\zeta'(0|L_{N-1})+U_0(\varepsilon)
\:,\label{Eodd}\eeq
where ${\cal U}_{N-2n}^\beta$ indicates the free energy density for a
scalar field with (negative) square mass $-\varrho_{N-2n}^2$
on an ultrastatic manifold with hyperbolic $H^{N-2n}$ spatial section
and $U_0(\varepsilon)$ is the vacuum energy.
With regard to the internal energy, we have at disposal a
renormalization procedure, which is well understood for $D=4$. In
fact, in Rindler and black hole space-times it is known that the
renormalized stress-energy tensor is finite at the horizon in the
Hartle-Hawking state \cite{scia81-30-327,brow85-31-2514},
corresponding to the temperature $\beta=\beta_H$. This is equivalent to
write
\beq
U(\beta)^{ren}=U_\varepsilon(\beta)-U_\varepsilon(\beta_H)+\mbox{finite part}\, ,
\label{uren}
\eeq
where $U_\varepsilon(\beta)$ is the divergent part of the internal energy and it
may be read off the Eqs. (\ref{Eeven}) and (\ref{Eodd}).
Thus, the divergences are present in the expression of the
renormalized internal energy, but only for some particular value of
$\beta$, say $\beta_H$ ($\beta_U$).
For example, in the case of Rindler space-time, such a value is
$\beta_U=2\pi a^{-1}$, the Unruh temperature (here $a$ is the acceleration),
while in the 4-dimensional black hole
background one has $\beta_H=8\pi MG$, the Hawking temperature.
In the general case, we may use of the same renormalization
procedure. Note, however, that the corresponding renormalized partition
function, free energy and entropy
remain divergent also at the distinguised temperature $\beta=\beta_H$.
\s{Some physical applications}
\label{S:PA}
As simple physical applications of the general formulae derived in
Sec.~\ref{S:FE}, here we consider the cases in which $\M^{N-1}$
is a homogeneous manifold with constant scalar curvature $\kappa$.
Of course we have the three possibilities
$\M^{N-1}\equiv\R^{N-1}$ ($\kappa=0$), $\M^{N-1}\equiv S^{N-1}$
($\kappa>0$) and finally $\M^{N-1}\equiv H^{N-1}$ ($\kappa<0$), but here we
only consider in more detail the first two cases.
The first one corresponds to the conformal treatment of
the $D$-dimensional Rindler space-time,
while the second appears when one studies the physics of black holes
near the horizon. For $D=4$, this case has been studied in
Ref.~\cite{cogn95-12-1927}).
\ss{Statistical mechanics for massive fields
in the Rindler $D$-dimensional space-time}
As we have already observed, after a conformal transformation, the spatial
section of the
Rindler space-time is of the kind condidered in the paper.
For this special case, the curvature of $\M^{N-1}$ is vanishing ($k=0$)
and so one easily has
\beq
K_{2n}(L_{N-1})=\frac{(-m^2)^n}{n!}
\frac{V_{N-1}}{(4\pi)^{\frac{N-1}2}}
\:,\hs
\zeta(z|L_{N-1})=\frac{V_{N-1}\Gamma(z-\frac{N-1}2)}
{(4\pi)^{\frac{N-1}2}\Gamma(z)}\,m^{N-1-2z}
\:,\label{0ok}\eeq
where $C=m^2$ has been put since Rindler is a flat manifold.
Now, using Eqs.~(\ref{FEeven}) and (\ref{FEodd}) together with
the two equations above, we obtain
\beq
F_\beta^{Rind}&=&\sum_{n=0}^{\frac{N-3}2}
\frac{V_{N-1}\,\varepsilon^{-(N-1-2n)}}{(N-1-2n)\,n!}
\at-\frac{m^2}{4\pi}\ct^n\,{\cal F}_{N-2n}^\beta
\nn\\&&\hs
-\frac{\pi}{12\beta^2}
\frac{V_{N-1}}{\Gamma(\frac{N+1}2)}
\aq\gamma+\psi(\fr{N+1}2)-\ln\fr{m^2\varepsilon^2}4\cq
\at-\frac{m^2}{4\pi}\ct^{\frac{N-1}2}
\nn\\&&
+\frac{V_{N-1}}{\Gamma(\frac{N+1}2)}
\at-\frac{m^2}{4\pi}\ct^{\frac{N-1}2}
\aq-\frac{\ln\beta}{4\beta}
+\frac1{2\pi\beta}
\int_0^{\ii}[\psi(ir)+\psi(-ir)][1-e^{-\beta r}]\,dr
\cq\:,\label{6.2}\eeq
\beq
F_\beta^{Rind}&=&\sum_{n=0}^{\frac{N-2}2}
\frac{V_{N-1}\,\varepsilon^{-(N-1-2n)}}{(N-1-2n)\,n!}
\at-\frac{m^2}{4\pi}\ct^n\,{\cal F}_{N-2n}^\beta
\nn\\&&\hs\hs
-\frac{\pi}{12\beta^2}
V_{N-1}\Gamma(-\fr{N-1}2)
\at-\frac{m^2}{4\pi}\ct^{\frac{N-1}2}
\:,\eeq
valid for even and odd $D$-dimension respectively.
In Eq.~(\ref{6.2}) $\gamma$ is the Euler constant.
The functions
${\cal F}_{N-2n}^\beta$ can be computed using Eqs.~(\ref{A3}),
(\ref{Ab3}) and (\ref{Ac3}) in the Appendix.
For example, when $D=4$, using Eq.~(\ref{Ab3-0}), the result is
\beq
F_\beta^{Rind}&=&-\frac{A\pi^2}{180\beta^4\varepsilon^2}
+\frac{Am^2(1-\ln\fr{m^2\varepsilon^2}4)}{48\beta^2}\nn\\
&&\hs+\frac{Am^2}{4\pi}
\aq\frac{\ln\beta}{4\beta}
-\frac1{2\pi\beta}
\int_0^{\ii}[\psi(ir)+\psi(-ir)][1-e^{-\beta r}]\,dr
\cq
\:,\label{erto}\eeq
where the transverse area $A=V_2$ has been introduced to compare the
latter formula with well known results (see for example
Refs. \cite{suss94-50-2700,kaba94-329-46,kaba95u-06,more95u-5}).
There is agreement in the massless case, but not in the massive case,
where we also obtain a finite contribution.
We conclude this section with some remarks on renormalization.
As we have seen above, in our formalism,
massive scalar fields in Rindler space-time can be easily treated,
because the optical spatial section turns out to be the hyperbolic
space $H^3$ and the harmonic analysis on such a manifold is well known.
The formulae are particularly simple in the massless case.
For example, in 4-dimensions, the total free energy
may be chosen in the form
\beq
F^{ren}(\beta)=-\frac{A}{45(8\pi)^2\varepsilon^2}
\aq\at\frac{\beta_U}{\beta}\ct^4+3\cq
\:,\eeq
where $\beta_U=2\pi$ is the Unruh temperature ($a=1$).
As a consequence, the entropy turns out to be
\beq S_\beta=\frac{8\pi^2 A}{45\varepsilon^2\beta^3}
\eeq
and it diverges for every finite $\beta$, but is zero at zero temperature
(the Fulling-Rindler state), which is correct, since we are dealing with a
pure state. At $\beta=\beta_U$, corresponding to the
Minkowski vacuum, we have a divergent entropy proportional to the
area, regardless of the fact that the Minkowski vacuum is a pure
state. This is also to be expected, since an uniformly accelerated
observer cannot observe the whole Minkowski space-time.
Finally with this renormalization prescription,
the internal energy should read
\beq
U^{ren}(\beta)=\frac{A}{15(8\pi)^2\varepsilon^2}
\aq\at\frac{\beta_U}{\beta}\ct^4-1\cq
\eeq
and this is vanishing and a
fortiori finite at $\beta=\beta_U$, as it should be. Furthermore, at
$\beta=\ii$, namely in the Fulling-Rindler vacuum, it is in agreement
with the result obtained in Ref.~\cite{brow86-33-2840}.
\ss{Statistical mechanics for massive fields
in a $D$-dimensional black hole background}
\label{S:BH}
Here we consider in more detail the case in which $\M^{N-1}=S^{N-1}$.
To justify this choice from a physical view point,
first of all we show that, near the horizon, a $D$-dimensional black hole
may be approximated by a manifold of this kind and so,
the thermodynamics can be derived by using the formulae of Sec.~\ref{S:FE}.
The static metric describing a $D$-dimensional Schwarzschild black hole
(we assume $D>3$) reads \cite{call88-311-673}
\beq
ds^2=-\aq1-\at\frac{r_H}r\ct^{D-3}\cq\,dx_0^2+
\aq1-\at\frac{r_H}r\ct^{D-3}\cq^{-1}\,dr^2
+r^2\,d\Omega_{D-2}
\:,\label{bh}
\eeq
where we are using polar coordinates, $r$ being the radial one and
$d\Omega_{D-2}$ the $D-2$-dimensional spherical unit metric.
The horizon radius is given by
\beq
r_H=\hat{G}_D M^{\frac1{D-3}}\:,
\hs\hat G_D=\aq\frac{2\pi^{\frac{D-3}2}\,G_D}
{(D-2)\Gamma(\frac{D-1}2)}\cq^{\frac1{D-3}}
\:,\label{rh}
\eeq
$M$ being the mass of the black hole and $G_D$
the generalized Newton constant. The associated Hawking temperature
reads $\beta_H=4\pi r_H/(D-3)$. The corresponding Bekenstein-Hawking
entropy may be computed by making use of
\beq
\beta_H=\frac{\partial S_H}{\partial M}\,.
\eeq
Thus we have
\beq
S_H=4\pi \fr{\hat{G}_D}{D-2} M^{\fr{D-2}{D-3}}
\eeq
{}From now on, we put $r_H=1$.
It may be convenient to redefine the radial Schwarzschild coordinate
$r=r(\rho)$ by means of the implicit relation
\beq
\rho^2&=&\frac4{D-3}\aq e^{r-1}
\exp\int\frac{dr}{r^{D-3}-1}\cq^\frac1{D-3}\:,\nn \\
&\sim&\frac2{D-3}(r-1)e^{\frac{(D-2)(r-1)}2}+\dots
\:,\label{ro}\eeq
and time $x_0=x'_0/b$, $b=(D-3)/2$ in order to have
$g_{00}=\rho^2+O(\rho^4)$.
In the new set of coordinates we have
\beq
ds^2=-\frac{1-r^{3-D}(\rho)}{b^2}\,dx_0'^2+
\frac{1-r^{3-D}(\rho)}{b^2\rho^2}\,d\rho^2
+r^2(\rho)\,d\Omega_{D-2}
\:,\eeq
and finally the optical metric reads
\beq
d\bar s^2 =-dx_0'^2+\frac{1}{\rho^2}\aq
d\rho^2+G(\rho)\,d\Omega_{D-2}\cq
\:,\label{OMbh}\eeq
where we have set
\beq
G(\rho)=\frac{(b\,r\,\rho)^2}{1-r^{3-D}}
=1+O(\rho^2)
\:.\eeq
{}From the latter equation we see that, near the horizon $\rho=0$,
we can set $G(\rho)=1$ and so the optical metric assumes the form
considered in previous Sections. In this approximation the manifold
$\M^{N-1}$ becomes the unit shpere $S^{N-1}$. We have
\beq
d\bar s^2 \simeq-dx_0'^2+\frac1{\rho^2}\aq
d\rho^2+d\Omega_{D-2}\cq
\:.\label{AMbh}\eeq
Such a metric can be considered as an approximation of the one of the
black hole in Eq.~(\ref{OMbh}) in the sense that, near the horizon, the
geodesics are essentially the same for both the metrics.
The metric (\ref{AMbh}) can be related to a manifold with curvature
$R^{\bar g}=-(D-1)(D-2)+O(\rho^2)$, then, according to Eq.~(\ref{Aq}),
the relevant operator becomes
\beq
\bar L_N=-\lap_N^{\bar g}-\varrho_N^2+C\rho^2 +O(\rho^4)
\:,\label{ROBH}\eeq
where now $C$ is a positive constant, which takes into account of
mass and curvature contributions to this order.
Note that since for the original manifold $R^g=0$, $\xi$ does not appear
in the formulae. This effectively happens if we approximate the metric
after the optical transformation has been done. More simply,
one can put $\xi=\xi_D$ in Eq.~(\ref{Aq}).
The discussion for arbitrary $D=N+1$ is quite involved even though it may
be done, since the $\zeta$-functions of the Laplace-Beltrami operators on
$S^{N-1}$ are known (see the Appendix).
As a more explicit example, now we consider a scalar field in
a 4-dimensional Schwarzschild background.
Using these techniques, such a case has been considered in
Ref.~\cite{cogn95-12-1927}, where we refer the interested reader for more
details.
We have $r_H=2MG$, $b=1/2$,
\beq
\rho=2(r-1)^{\frac12}e^{(r-1)/2}
\:,\label{ro1}\eeq
and
\beq
r=1+\frac{\rho^2}{4}-\frac{\rho^4}{16}+O(\rho^6)
\:.\label{456}\eeq
Then, according to Eq.~(\ref{ROBH}), the relevant operator becomes
\beq
\bar L_3=-\bar\lap_3-1+C\rho^2
\:,\eeq
where $C=m^2+1/3$ takes into account of the curvature
$R^{\bar g}=-6+2\rho^2$ of the optical manifold.
Now, directly using Eqs.~(\ref{logPF-Jacobi}), (\ref{FE}),
(\ref{FEeven}) and (\ref{Ab3-0}), after
the replacement $\beta\to\beta/2$ due to the redefinition of the
Schwarzschild time (remember that $b=1/2$), for the total free energy
we obtain
\beq
F^{bh}(\beta) &=&-Aj_\varepsilon+\frac14\zeta(-\fr12|\bar L_3)
-\frac{2\pi^2A}{45\varepsilon^2\beta^4}-\frac{A}{12\beta^2}\aq
\frac{\zeta'(0|L_2)}2-\zeta(0|L_2)\ln\frac\ep2\cq \nn\\&&
-\frac{A\zeta(0|L_2)}{16\pi\beta}\ln\frac\be2
+\frac{A\zeta(0|L_2)}{8\pi^2\beta}\int_0^\ii
\ln\at1-e^{-\beta r/2}\ct\, \aq\psi(ir)+\psi(-ir)\cq\,dr
\:,\label{bhf} \eeq
where we have written the Jacobian contribution to the partition
function due to the conformal transformation in the form $A\beta j_\varepsilon$,
and now $A=4\pi r_H^2$ is the transverse area of the black hole.
The $\zeta$-function related to the operator $L_2$ on the sphere is
given in the Appendix by Eq.~(\ref{99}) with $C=m^2+1/3$,
then $\zeta(0|L_2)=m^2$.
The leading divergence, due to the optical volume,
is proportional to the horizon area \cite{thoo85-256-727},
but in contrast with the Rindler case, a
logarithmic divergence is also present, similar to the one found in
Refs.~\cite{solo94u-246,deal94u-347}.
This is a feature of even dimensions (see Sec.~\ref{S:FE}).
Let us briefly discuss the renormalization of the internal
energy in this particular case.
We recall that one needs a renormalization in order to remove the
vacuum divergences. These divergenges, as well as the
Jacobian conformal factor, do not contribute to the entropy.
However the situation presented here is complicated
by the presence of horizon divergences, controlled by the cutoff parameter
$\varepsilon$. In the 4-dimensional Schwarzschild space-time, it is known that the
renormalized stress-energy tensor is well defined at the horizon in
the Hartle-Hawking state \cite{scia81-30-327,brow85-31-2514}, which in
our formalism corresponds to the Hawking temperature $\beta=\beta_H$.
The renormalized internal energy reads (the dots stay for finite
contributions at the horizon, which we do not write down because their
value depend on the approximation made)
\beq
U^{ren}(\beta)&=&\frac{A}{30(8\pi)^2\varepsilon^2}
\aq\at\frac{\beta_H}{\beta}\ct^4-1\cq
-\frac{A}{3(8\pi)^2}\ln\varepsilon \aq\at\frac{\beta_H}{\beta}\ct^2-1\cq
\:\:+\dots \:,\eeq
which has no divergences for $\beta=\beta_H$, the Hawking temperature,
while the entropy
\beq
S_\beta=\frac{8\pi^2A}{45\varepsilon^2\beta^3}
-\frac{A\ln\varepsilon}{6\beta}\:\:+\dots\:,\label{S1}\eeq
also for $\beta=\beta_H$ contains the well known divergent
term proportional to the horizon
area \cite{thoo85-256-727} and, according to Ref.~\cite{solo94u-246},
a logarithmic divergence too. Eq.~(\ref{S}) is vanishing in the
Boulware vacuum corresponding to $\beta=\ii$.
{}From this renormalization procedure we get for the renormalized black hole
free energy
\beq
F^{ren}(\beta) &=&-\frac{A}{90(8\pi)^2\varepsilon^2}
\aq\at\frac{\beta_H}{\beta}\ct^4+3\cq
+\frac{A}{3(8\pi)^2}\ln\varepsilon
\aq\at\frac{\beta_H}{\beta}\ct^2+1\cq\nn\\
&&\hs -\frac{A}{12\beta^2}\aq
\frac{\zeta'(0|L_2)}2+m^2\ln2\cq
-\frac{Am^2}{16\pi\beta}\ln\frac\be2\nn\\
&&\hs\hs +\frac{Am^2}{8\pi^2\beta}
\int_0^\ii\ln\at1-e^{-\beta r/2}\ct\,
\aq\psi(ir)+\psi(-ir)\cq\,dr
\:.\label{bhf2} \eeq
In the general case, the discussion is quite similar and it can be
performed by using Eqs.~(\ref{FEeven}) or (\ref{FEodd})
with the replacement $\beta\to(D-3)\beta/2$.
\s{Conclusions}
\label{S:C}
In this paper the first quantum corrections to the thermodynamic
quantities of fields in a
$D$-dimensional Rindler-like space have been investigated
making use of conformal transformation techniques and $\zeta$-function
regularization.
In this way, we have worked within the so
called optical manifold, which is ultrastatic, and the
use of finite temperature methods is quite straightforward.
The general form of the
horizon divercences of the free energy has been obtained as a function of
free energy densities of fields having negative square masses (absence
of the gap in the Laplace operator spectrum) on ultrastatic manifolds
with hyperbolic spatial section $H^{N-2n}$ and of the Seeley-DeWitt
coefficients $K_{2n}(L_{N-1})$ of the Laplace operator on $\M^{N-1}$.
Since there exists recurrence relations for free energy densities
(see the Appendix), it is sufficient to study the cases $D=3$ and $D=4$
($D=4$ and $D=5$ for applications to black holes).
The leading divergence can be seen to be given by
the volume of the spatial section of the optical manifold.
A finite contribution is
also obtained and this depends on $\zeta(0|L_{N-1})$ and on its first
derivative. For $D=4$, our results are consistent with the ones obtained
with brick wall, path-integral and canonical methods
\cite{ghos94-73-2521,deal95u-33,barb95u-155}.
With regard to physical applications, we have used the general results
on finite temperature field theory
in order to investigate the quantum corrections to the
Bekenstein-Hawking entropy for massive fields in a
large mass black hole background.
This approach gives rise to a leading divergence for the entropy
similar to the one obtained for the Rindler case background,
but in this case other divergent contributions are present
and their structure depend on the dimension of the
space-time considered.
Here we have shown how it is possible to get the general form valid for
an arbitrary dimension and we have explicitly considered the case
$D=4$.
We also would like to mention the results obtained in
Ref.~\cite{frol93-48-4545}, where the contributions to the $4$-dimensional
black hole entropy due to modes located inside and near the horizon have been
evaluated using a new invariant statistical mechanical definition for
the black hole entropy. The finite contributions, namely the ones
indipendent on the horizon cutoff, are compatible with our results.
As far as the horizon divergences are concerned we recall that they may
be interpreted physically in terms of the infinite gravitational redshift
existing between the spatial infinity, where one measures the generic
equilibrium temperature and the horizon, which is classically
unaccessible for the Schwarzschild external observer.
Furthermore, we have argued that they are absent in the
internal energy at the Unrhu-Hawking temperature. However, they remain in
the entropy and in the other thermodynamical quantities, as soon as
one assumes the validity of the usual thermodynamical relations. For
$D=4$, a possible way to deal with such divergences has been suggested in
Refs.~\cite{thoo85-256-727,frol93-48-4545,frol94-74-3319}, where it has
been argued that the quantum fluctuations at the horizon might provide
a natural cutoff. In particular, choosing the horizon cutoff parameter
of the order of the Planck length ($\varepsilon^2\sim G$), the leading "divergence",
evaluated at the Hawking temperature, turns out to be of the form of
the the "classical" Bekenstein-Hawking entropy. This seems a
reasonable assumption, because we have worked within the fixed
background approximation. However one should
remark that other terms are present, giving contributions which
violate the area law. A more elaborate discussion for $D=4$ can be
found in Ref.~\cite{frol94u-211}.
Alternatively, one may try to relate the horizon
divergences to the ultraviolet divergences of quantum gravity, thus
arriving at the theory of superstring progagating in a curved
space-time \cite{suss94-50-2700} or at the renormalization group approach
\cite{odin95u-27}.
Finally, we mention that there has been the proposal to absorbe
the horizon divergences, at least for $D=2,3,4$, by making use of
the standard ultraviolet gravitational constant renormalization
\cite{suss94-50-2700,solo95-51-609,deme95u-3,furs94u-20,lars95u-66}.
This proposal is essentially based on the use of Euclidean section
with a conical singularity and
associated heat kernel expansion. The problematic issue consisting in
dealing with a finite temperature theory in a non ultrastatic space-time
is solved working within a non vanishing
conical singularity and interpreting the deficit angle of the
Euclidean compactified time as the inverse of the
equilibrium temperature (the absence of the conical singularity gives
the correct Hawking temperature). However, the resulting partition function
has, apparently, a wrong dependence on this "temperature".
Furthermore, it seems that the
only divergences present are the usual ultraviolet ones
associated with the definition of the partition function as determinant
of an elliptic operator. These divergences are then absorbed in the
gravitational constant renormalization. However, the naive use of
$\zeta$-function regularization should get rid off these ultraviolet
divergences.
Thus it seems to exist a disagreement between this approach and our
approach based on the conformal transformation techniques. It is our
opinion that this disagreement might depend on a non commutative
property present in the evaluation of heat kernel trace on a cone. It
should be interesting to elucidate this issue.
\ack{We wish to thank E.S. Moreira Jnr. for pointing out a mistake
in Eq. (\ref{erto}) in the first version of the manuscript
and L. Vanzo for discussions.}
|
\chapter{Introduction.}
The study of topological quantum field theories
\REF\wittop{E. Witten \journalComm. Math. Phys.&117(88)353.}
\REF\bbrt{For a review see D. Birmingham, M. Blau, M. Rakowski and G. Thompson
\journalPhys. Rep.&209(91)129.}
[\wittop,\bbrt] has aroused great interest recently.
These theories, being endowed with a BRST symmetry, are
models with no local degrees of freedom, so
all local excitations can be eliminated once the
topological symmetry has been fixed. In two dimensions
there exist conformal field theories which, in addition, are
topological field theories. These are the so-called Topological
Conformal Field Theories (TCFT)
\REF\dij{R. Dijkgraaf, E. Verlinde and H. Verlinde
\journalNucl. Phys.&B352(91)59.;``Notes on topological string theory
and 2d quantum gravity", Proceedings of the Trieste spring
school 1990, edited by M.Green et al. (World Scientific,
Singapore,1991).}[\dij].
A method to generate new TCFT's
consists in studying the BRST structure of different chiral algebras
that extend the Virasoro symmetry. This method has been applied
in ref.
\REF\tca{J. M. Isidro and A. V. Ramallo\journalPhys. Lett.&B316(93)488.}
[\tca] to the case of an affine Lie algebra, whereas in ref.
\REF\tkt{J. M. Isidro and A. V. Ramallo\journalPhys. Lett.&B340(94)48.}
[\tkt] this analysis was extended to the case of a superconformal
current algebra. It is the purpose of this paper to explore, using
the same procedure, the TCFT's based on affine Lie
superalgebras.
The topological symmetry of a TCFT is encoded in its topological
algebra, which is the operator algebra closed by the chiral
algebra of the TCFT and the BRST current. It was checked in refs.
\REF\gln{J. M. Isidro and A. V. Ramallo\journalNucl. Phys.&B414(94)715.}
[\tca,\tkt,\gln] that the topological algebra of a TCFT
possessing a non-abelian current algebra must include operators of
dimensions one, two and three. This algebra is the so-called
Kazama algebra
\REF\kazama{Y. Kazama \journalMod. Phys. Lett.&A6(91)1321.}[\kazama], which
differs from the standard twisted $N=2$ superconformal algebra
\REF\EY{T. Eguchi and S.-K. Yang \journalMod. Phys. Lett.&A4(90)1653;
T. Eguchi, S. Hosono and S.-K. Yang \journalComm. Math. Phys.&140(91)159.}[\EY].
The former includes two dimension-three operators and
can be regarded as an extension of the latter. The extended
nature of the Kazama algebra seems to be an unavoidable
consequence of the underlying non-abelian current symmetry.
The representation of the BRST symmetry found in refs.
[\tca,\tkt,\gln] requires the level of the current algebra to
be fixed to some critical value related to the dual Coxeter
number of the Lie algebra. When the matter sector of
the currents is realised by means of two decoupled currents, it
is only the sum of the two levels that is constrained. In this
two-current realization, the TCFT for a Lie algebra ${\cal G}$ has a
nice interpretation as a ${\cal G}/{\cal G}$ coset model
\REF\witGG{E. Witten \journalComm. Math. Phys.&144(92)189.}
[\witGG]. After a suitable
deformation these theories have been shown to be related to
non-critical $W$-strings
\REF\yank{M. Spiegelglas and S. Yankielowicz
\journalNucl. Phys.&393(93)301; O. Aharony et al.\journalNucl. Phys.&B399(93)527
\journalPhys. Lett.&B289(92)309 \journalPhys. Lett.&B305(93)35.}
\REF\hu{H.L. Hu and M. Yu \journalPhys. Lett.&B289(92)302
\journalNucl. Phys.&B391(93)389.}
\REF\sadov{V. Sadov\journalInt. J. Mod. Phys.&A8(93)5115.}
[\yank,\hu,\sadov].
Lie superalgebras seem to play an important role in the
construction and classification of extended superconformal
algebras
\REF\sevrin{A. Sevrin, W. Troost and A. van Proeyen
\journalPhys. Lett.&B208(88)447.}
\REF\petersen{K. Ito, J. O . Madsen and J. L. Petersen
\journalPhys. Lett.&B318(93)315.}
[\sevrin,\petersen] and supersymmetric Toda field theories.
Given a Lie superalgebra which admits
a purely fermionic simple root system, one can
construct an $N=1$ supersymmetric Toda model
\REF\komata{S. Komata, K. Mohri and H. Nohara
\journalNucl. Phys.&B359(91)168.}
\REF\evans{J. Evans and T. Hollowood\journalNucl. Phys.&B352(91)723.}
\REF\inami{T. Inami and K.-I. Izawa\journalPhys. Lett.&B225(91)523.}
[\komata,\evans,\inami]. Application of the method of
hamiltonian reduction to Lie (super)algebras
leads to (super) ${\cal W}$-algebras
\REF\bershadsky{M. Bershadsky and H. Ooguri
\journalComm. Math. Phys.&126(89)49\journalPhys. Lett.&B229(89)374.}
\REF\schoutens{For a review see P. Bouwknegt
and K. Schoutens\journalPhys. Rep.&223(93)183.}
[\bershadsky,\schoutens].
There is also a possibility that string theories
may be classified in terms of superalgebras. For
example, the twisted superconformal symmetry for
strings with $N-2$ supersymmetries has been
constructed in
\REF\boresch{A. Boresch, K. Landsteiner, W. Lerche and A. Sevrin,
\journalNucl. Phys.&B436(95)609.}
[\boresch] via the quantum
hamiltonian reduction of ${\rm osp}(N\vert 2)$.
It therefore seems desirable to extend the
formalism initiated in [\tca,\tkt,\gln] to cover the
general case of an arbitrary Lie superalgebra. We shall show
below that the approach followed in refs. [\tca,\tkt] can be
easily generalised to TCFT's possessing an affine Lie
superalgebra. The BRST algebra for these theories is the same as
in the bosonic case, which yields a new representation of the
topological Kazama algebra. Moreover we provide arguments to
support the idea that these theories are related to non-critical
superstring theories.
This paper is organised as follows. In section 2, after
a brief introduction to current
superalgebras and the associated Sugawara constructions
in two dimensions, the
topological algebra is developed explicitly. A two-current
construction is also possible, which
paves the way for the interpretation of the theory as a gauged,
supergroup-valued
Wess-Zumino-Witten (WZW) model; this is done in section 3.
In section 4 we turn our attention to the relation
of the model constructed with the non-critical superstring
theories. We shall find a suggestive connection between the
${\rm osp}(1\vert 2)$ and ${\rm osp}(2\vert 2)$ theories and
non-critical superstring theories with one and two
supersymmetries respectively. These results generalise the
relation between non-critical $W_N$-strings and ${\rm sl}(N)/{\rm sl}(N)$
coset theories [\yank,\hu,\sadov]. Finally, our work is
summarised in section 5, together with some conclusions,
comments and suggestions for future work.
\chapter{Construction of the topological algebra.}
Let ${\cal G}$ be a finite-dimensional superalgebra over
the complex field
\REF\kac{V. G. Ka\v c\journalComm. Math. Phys.&53(77)31 \journalAdv. in Math.&26(77)8.}
\REF\scheunert{M. Scheunert,
``The Theory of Lie Superalgebras",
{\sl Lect. Notes in Math.} 716,
Springer-Verlag, Berlin (1979).}
\REF\frappat{L. Frappat, A. Sciarrino and P. Sorba
\journalComm. Math. Phys.&121(89)457.}
[\kac,\scheunert,\frappat]. Call
${\cal G}_B$ and ${\cal G}_F$ the even
(\ie, bosonic) and odd (\ie, fermionic)
subspaces of ${\cal G}$, spanned respectively by basis vectors
$T_a$, $a=1,2, \ldots d_B$, and $T_{\alpha}$, $\alpha =1,2,
\ldots d_F$, where $d_B$
and $d_F$ are the respective dimensions of
${\cal G}_B$ and ${\cal G}_F$.
In the
following, latin indices $a, b,\ldots$ will run from
1 to $d_B$, while greek
indices $\alpha, \beta,\ldots$ from 1 to $d_F$.
The {\it superdimension}
of ${\cal G}$
is $d_{B}-d_{F}$. Denote the (generalised) Lie bracket by $[\,,]$. One has
$$
[T_a,T_b]=f_{ab}^{c}T_c\qquad
[T_a,T_{\alpha}]=f_{a\alpha}^{\beta}T_{\beta}\qquad
[T_{\alpha},T_{\beta}]=f_{\alpha \beta}^c T_c.
\eqn\za
$$
The generalised Lie bracket satisfies a graded Jacobi identity
$$
(-1)^{g(x)g(z)}[x,[y,z]] + (-1)^{g(y)g(x)}[y,[z,x]]
+ (-1)^{g(z)g(y)}[z,[x,y]] =0
\eqn\zb
$$
for any $x$, $y$ and $z$ in ${\cal G}$,
where $g(x)=0$ if $x\in {\cal G}_B$
and $g(x)=1$ if $x\in {\cal G}_F$.
${\cal G}$ will be assumed to possess
a real, non-degenerate, supersymmetric
bilinear form $(\,,)$ such that ${\cal G}_B$
and ${\cal G}_F$ are orthogonal. It will also be
assumed to satisfy the {\it invariance
property}
$$
([x,y],z)=(x,[y,z])
\eqn\zc
$$
for all $x$, $y$ and $z$ in ${\cal G}$. Call
$g_{ab}=(T_a,T_b)$, $ g_{\alpha\beta}=(T_{\alpha},T_{\beta})$;
one has
$g_{ab}=g_{ba}$, $g_{\alpha\beta}=-g_{\beta\alpha}$.
Indices are raised and
lowered by contraction with the metric
tensor and its inverse according to
$T^a=g^{ab}T_b$, $T_a=g_{ab}T^b$, $T^{\alpha}=T_{\beta}g^{\beta\alpha}$,
$T_{\alpha}=T^{\beta}g_{\beta\alpha}$. Upon lowering of its upper index,
$$
f_{abc}=f_{ab}^{d}g_{dc},\quad
f_{a\beta\gamma}=f_{a\beta}^{\mu}g_{\mu\gamma},\quad
f_{\alpha\beta c}=f_{\alpha\beta}^{d}g_{dc},
\eqn\zd
$$
the structure constants become superantisymmetric. The Jacobi identity
(eq. \zb) imposes additional conditions on the structure constants.
Some of these conditions are:
$$
\eqalign{
&f_{\alpha\beta}^{c}\,f_{ca}^{d}\,-\,
f_{\beta a}^{\lambda}\,f_{\lambda \alpha}^{d}\,+\,
f_{ a\alpha}^{\lambda}\,f_{\lambda \beta}^{d}\,=\,0\cr
&f_{ab}^{c}\,f_{c\alpha}^{\beta}\,+\,
f_{b\alpha}^{\lambda}\,f_{\lambda a}^{\beta}\,+\,
f_{ \alpha a}^{\lambda}\,f_{\lambda b}^{\beta}\,=\,0\cr
&f_{\alpha\beta}^{c}\,f_{c\gamma}^{\delta}\,+\,
f_{\beta \gamma}^{c}\,f_{c \alpha}^{\delta}\,+\,
f_{ \gamma\alpha}^{c}\,f_{c \beta}^{\delta}\,=\,0\cr}
\eqn\zh
$$
Other relations satisfied by the structure constants can be
obtained from the quadratic Casimir operator of the
superalgebra. If $C_A$ denotes the value of this operator in
the adjoint representation, one has:
$$
\eqalign{
&g^{bc} f_{ab}^{d} f_{dc}^{e}\,+\,
g^{\alpha\beta} f_{a\alpha}^{\gamma}
f_{\gamma\beta}^{e}
=C_A\,\delta ^{e}_{a}\cr\cr
&g^{bc} f_{\alpha b}^{\gamma} f_{\gamma c}^{\mu}\, =\,
g^{\beta\gamma} f_{\alpha\beta}^{c} f_{c\gamma}^{\mu}\,=\,
{C_A\over 2}\,\, \delta ^{\mu}_{\alpha}\cr}
\eqn\zi
$$
A conformal current superalgebra is generated
by a set of holomorphic
bosonic $J_a(z)$ and fermionic currents $J_{\alpha}(z)$,
satisfying the following operator product expansions (OPE's):
$$
\eqalign{
J_a(z)J_b(w)=&{k g_{ab}\over (z-w)^2} +
{f_{ab}^{c}\over z-w}J_c(w)\cr
J_a(z)J_{\beta}(w)=&{f_{a\beta}^{\gamma}\over z-w}J_{\gamma}(w)\cr
J_{\alpha}(z)J_{\beta}(w)=&{k g_{\alpha\beta}\over (z-w)^2} +
{f_{\alpha\beta}^{c}\over z-w}J_c(w).
\cr}
\eqn\ze
$$
$k$ is the level of the current superalgebra.
The currents $J_a$ and $J_{\alpha}$ can
be made into Virasoro primary fields by
application of the Sugawara
construction, whereby a bilinear in the $J$'s
$$
T^J=N(g^{ab}J_{a}J_{b} + g^{\alpha\beta}J_{\alpha}J_{\beta})
\eqn\zf
$$
is required to satisfy a Virasoro algebra,
such that all the currents $J$ have
conformal dimension 1 with respect to $T^J$;
the normalisation constant $N$ is fixed precisely by this
requirement. This analysis
is standard and has been carried out,
for the case of a Lie superalgebra,
\REF\goddard{P. Goddard, D. Olive and
G. Waterson\journalComm. Math. Phys.&112(87)591.}
\REF\jarvis{P. D. Jarvis and R. B. Zhang\journalPhys. Lett.&B215(88)695.}
\REF\henningson{M. Henningson\journalInt. J. Mod. Phys.&A6(91)1137.}
\REF\troost{A. Deckmyn and W. Troost\journalNucl. Phys.&B370(92)231.}
\REF\fujitsu{A. Fujitsu\journalMod. Phys. Lett.&A8(93)1763.}
in [\goddard,\jarvis,\henningson,\troost,\fujitsu]. The
value of $N$ is found to be
$$
N={1\over 2k + C_A},
\eqn\zg
$$
and the operator
$$
T^J={1\over 2k + C_A}(g^{ab}J_{a}J_{b}
+ g^{\alpha\beta}J_{\alpha}J_{\beta})
\eqn\zj
$$
closes a Virasoro algebra with a central charge $c_J$ given
by $$
c_J= {2k (d_B - d_F)\over 2k+ C_A}.
\eqn\zk
$$
Let us now describe how one can construct a TCFT based on
the algebra \ze. The basic ingredient in our construction
is the introduction of a ghost sector such that one can
realise the BRST symmetry of the superalgebra \ze. We
shall use this BRST symmetry as the topological symmetry
of the TCFT. In general, in order to represent the BRST
symmetry of a given chiral algebra, one has to introduce
a fermionic(bosonic) ghost system for every
bosonic(fermionic) generator of the algebra. To these
ghost systems one must assign ghost numbers in such a way
that each antighost and its associated generator have the
same spin. According to these general rules we must
introduce in our case a spin-one ghost system for each
current of the superalgebra \ze. Moreover, the
topological symmetry we are trying to implement is such
that the BRST variation of the antighost equals the
corresponding total current.
Let us denote by $(\rho_a,\gamma^a)$ to the fermionic
ghosts for the currents $J_a$ whereas the bosonic fields
$(\lambda_{\alpha}, \eta^{\alpha})$ will correspond to
the currents $J_{\alpha}$. Let us choose our conventions
in such a way that the fermionic ghost fields $\rho _{a}$,
$\gamma ^{b}$ satisfy the OPE
$$
\rho_a(z) \gamma^b(w)= {-\delta^{b}_{a}\over z-w}.
\eqn\zl
$$
$\rho_a$ and $\gamma^b$ will be assumed to have
conformal weights 1 and 0,
and will be assigned ghost numbers $-1$ and $+1$,
respectively. Their
energy-momentum tensor $T^{(\rho\gamma)}$ is given by
$$
T^{(\rho\gamma)}=\rho_{a}\partial \gamma^{a},
\eqn\zm
$$
and has a central charge $c_{(\rho\gamma)}=-2d_B$.
Similarly the bosonic ghost fields $\eta ^{\alpha}$,
$\lambda _{\beta}$ will satisfy
$$
\eta^{\alpha}(z) \lambda_{\beta}(w)= {-\delta^{\alpha}_{\beta}\over z-w}.
\eqn\zn
$$
$\eta ^{\alpha}$ and $\lambda _{\beta}$ have conformal weights 0
and 1 and ghost numbers $+1$ and $-1$, respectively,
and their energy-momentum tensor is:
$$
T^{(\eta\lambda)}=\partial \eta^{\alpha} \lambda_{\alpha}.
\eqn\zo
$$
The central charge of $T^{(\eta\lambda)}$ is
$c_{(\eta\lambda)}=2d_F$. Altogether, the operator
$$
T=T^J + T^{(\rho\gamma)} + T^{(\eta\lambda)}
\eqn\zp
$$
closes a Virasoro algebra with a central charge given by
$$
c_{(tot)}=[\,{2k \over 2k+ C_A} -
2]\,(d_B - d_F).
\eqn\zq
$$
A TCFT is expected to have a vanishing Virasoro anomaly.
Notice that the condition $c_{(tot)}=0$ is satisfied when
$k=-C_A$. This value of the level determines the
topological point of the current + ghost system under
consideration. Another way to see how a
topological theory comes about is the following. Let us
consider the combination
$$
J_{a}^{{\rm gh}}=f_{ab}^{c}\gamma^{b}\rho_{c} -
f_{a\beta}^{\gamma}\eta^{\beta}\lambda_{\gamma},
\eqn\zr
$$
which is a bosonic, zero ghost-number field with
conformal weight 1 with
respect to $T$ in eq. \zp. Similarly, consider the object
$$
J_{\alpha}^{{\rm gh}}= f_{\alpha\mu}^{b}\rho_{b}\eta^{\mu} +
f_{\alpha b}^{\mu}\gamma^{b}\lambda_{\mu},
\eqn\zs
$$
with the same quantum numbers as above,
but fermionic statistics.
$J_{a}^{{\rm gh}}$ and $J_{\alpha}^{{\rm gh}}$
can be checked to verify the
algebra \ze, with a level $k=C_A$. Actually one can
easily verify that $J_{a}^{{\rm gh}}$ and
$J_{\alpha}^{{\rm gh}}$ represent the generators of the
superalgebra in the space of ghost fields. Therefore
the {\it total} bosonic and
fermionic currents read
$$
{\cal J}_{a}= J_{a}+J_{a}^{{\rm gh}}\qquad
{\cal J}_{\alpha}=J_{\alpha}+J_{\alpha}^{{\rm gh}},
\eqn\zt
$$
and satisfy the algebra \ze\ with a
total level $k_{(tot)}= k + C_A$.
According to the general arguments of refs. [\tca,
\tkt], a topological current superalgebra should now
appear at that particular value of $k$ for which
$k_{(tot)}$ vanishes, \ie, for
$k=-C_A$. That this is indeed
correct is confirmed by the fact that
this latter value, when substituted into
eq. \zq, gives $c_{(tot)}=0$.
We now work out the topological structure
present in the theory. To begin
with, a nilpotent BRST current $Q$ having
fermionic statistics, conformal
weight 1 and ghost number $+1$ is needed.
With the fields at hand, the
combination
$$
Q=-\gamma^a J_a - {1\over 2}f_{ab}^{c}\gamma^a\gamma^b\rho_c +
f_{a\beta}^{\mu}\gamma^a\eta^{\beta}\lambda_{\mu} - \eta^{\alpha} J_{\alpha}
- {1\over 2}f_{\alpha\beta}^{c}
\eta^{\alpha}\eta^{\beta}\rho_c
\eqn\zu
$$
satisfies the necessary requirements. Indeed $Q$ is the
canonical BRST charge for the Lie superalgebra \ze.
A tedious
although straightforward
calculation shows that
$$
Q(z)Q(w) = {k+C_A\over z-w}\,(g_{ab}\partial\gamma^a\gamma^b +
g_{\alpha\beta}\partial \eta^{\alpha}\eta^{\beta}),
\eqn\zv
$$
thus confirming again that only
for the critical value $k=-C_A$ is it
possible to have a nilpotent topological symmetry.
From now on we will always
assume that we are working at the critical level.
One can now suspect that all the operators present
in the theory appear in
BRST doublets. To prove that this statement is true,
let us begin with the
total Ka\v c--Moody currents as given in eq. \zt. We have
$$
\eqalign{
Q(z)\rho_a(w)=&{1\over z-w}{\cal J}_a(w)\cr
Q(z){\cal J}_a(w)=&0\cr
Q(z)\lambda_{\alpha}(w)=&{1\over z-w}{\cal J}_{\alpha}(w)\cr
Q(z){\cal J}_{\alpha}(w)=&0,\cr}
\eqn\zw
$$
so $(\rho_a,{\cal J}_a)$ and $(\lambda_{\alpha},{\cal J}_{\alpha})$
form weight 1 topological doublets. In eq. \zw\ one
notices that the BRST variations of the antighosts
$\rho_a$ and $\lambda_{\alpha}$ are the total currents ${\cal J}_a$
and ${\cal J}_{\alpha}$ respectively, which confirms the
correctness of our choice for $Q$. The operator algebra
of $\rho_a$, $\lambda_{\alpha}$, ${\cal J}_b$ and
${\cal J}_{\beta}$ closes as follows:
$$
\eqalign{
{\cal J}_a(z){\cal J}_b(w)=
&{f_{ab}^c\over z-w}{\cal J}_c(w)\cr
{\cal J}_a(z){\cal J}_{\beta}(w)=
&{f_{a\beta}^{\gamma}\over z-w}{\cal J}_{\gamma}(w)\cr
{\cal J}_{\alpha}(z){\cal J}_{\beta}(w)=
&{f_{\alpha\beta}^{c}\over z-w}{\cal J}_c(w)\cr
{\cal J}_a(z)\rho_b(w)=&{f_{ab}^c\over z-w}\rho_c(w)\cr
{\cal J}_a(z)\lambda_{\alpha}(w)=&{f_{a\alpha}^{\beta}\over z-w}\lambda_{\beta}(w)\cr
{\cal J}_{\alpha}(z)\rho_b(w)=&-{f_{\alpha b}^{\gamma}\over z-w}\lambda_{\gamma}(w)\cr
{\cal J}_{\alpha}(z)\lambda_{\beta}(w)=&-{f_{\alpha\beta}^c\over z-w}\rho_c(w).\cr}
\eqn\zx
$$
The topological character of the theory is ensured if the
energy-momentum tensor $T$ in eq. \zp\ is $Q$-exact. In
that case the BRST ancestor of $T$, denoted by $G$,
would be a weight 2 fermionic field with
ghost number $-1$. A glance at
eq. \zw\ can give some idea about
its expression: a Sugawara-like bilinear
of the form $g^{ab}\rho_a J_b$, $g^{\alpha\beta}\lambda_{\alpha}J_{\beta}$, with some
appropriate coefficients, will do. The precise combination
$$
G= {-1\over C_A} (g^{ab}\rho_a J_b + g^{\alpha\beta}\lambda_{\alpha}J_{\beta}),
\eqn\zz
$$
where the overall coefficient equals the one
in eq. \zj\ for the
critical level, satisfies
$$
Q(z)G(w)= {d_B-d_F\over (z-w)^3} +
{1\over (z-w)^2}R(w) + {1\over z-w}T(w).
\eqn\zaa
$$
In eq. \zaa, $R$ is a weight 1 bosonic
field with ghost number zero given by
$$
R=\rho_a\gamma^a + \lambda_{\alpha}\eta ^{\alpha}.
\eqn\zab
$$
Eq. \zaa\ is characteristic of two-dimensional
topological models. The
residue at the simple pole is the energy-momentum
tensor, which proves its
BRST-exactness, while the operator appearing
at the double pole is a
$U(1)$ current. Indeed, one easily checks that
$$
R(z)R(w)= {d_B-d_F\over (z-w)^2},
\eqn\zac
$$
and $R$ can be understood as the ghost-number
current. Indeed the $R$-charges of the different fields
coincide with the ghost numbers we have assigned them.
As for the triple pole
in eq. \zaa, the coefficient is a c-number
called the {\it topological
dimension} $d$ of the model, which now equals
the superdimension of the current
superalgebra. The same arguments as those
developed in [\tca] lead us to
conclude
that we are in fact describing a topological
sigma model for the underlying
supergroup manifold. A second BRST-doublet is
thus given by $(G,T)$, and the
OPE
$$
Q(z)R(w)={-1\over z-w}Q(w)
\eqn\zad
$$
shows that $(R,Q)$ too are BRST partners.
One can now compute the remaining
algebra between the generators above, with the result that
$$
\eqalign{
T(z)Q(w)=&{1\over (z-w)^2}Q(w) + {1\over z-w}\partial Q(w)\cr
T(z)R(w)=&{-(d_B-d_F)\over (z-w)^3} + {1\over (z-w)^2}R(w) +
{1\over z-w}
\partial R(w)\cr
T(z)G(w)=&{2\over (z-w)^2}G(w) + {1\over z-w}\partial G(w)\cr
R(z)G(w)=&{-1\over z-w}G(w).\cr}
\eqn\zae
$$
The above OPE's are exactly
those obtained upon twisting the $N=2$
superconformal algebra
\REF\LVW{W. Lerche, C. Vafa and N. P. Warner
\journalNucl. Phys.&B324(89)427.}
[\LVW,\EY] (the so-called {\it topological algebra}),
so one might be led to believe that such an
algebra is also present here. However, there is a
fundamental difference now, because the BRST partner of
$T$, $G$, is {\it not}
nilpotent. Instead one has
$$
G(z)G(w)={1\over z-w}W(w),
\eqn\zaf
$$
where $W$ is a bosonic, dimension 3 operator
with ghost number $-2$ given by
$$
\eqalign{
W=&-{1\over C_A}(\partial \rho_a\rho^a +
\partial \lambda_{\alpha}\lambda^{\alpha})\cr
+&{1\over (C_A)^2}\,(f_{ab}^c\rho^a \rho^b J_c +
f_{\alpha \beta}^c \lambda^{\alpha} \lambda^{\beta} J_c +
2 f_{a\beta}^{\gamma} \rho^a \lambda^{\beta} J_{\gamma} ).\cr}
\eqn\zag
$$
Since all operators in the theory so far
have appeared as BRST doublets, one
would expect this to hold for $W$, too.
And given that $W$ is BRST-closed, \ie,
$$
Q(z)W(w)=0,
\eqn\zah
$$
we must look for a BRST ancestor for $W$
with the following quantum numbers:
fermionic
statistics, conformal weight 3, and ghost
number $-3$. The combinations
$f_{abc}\rho^a \rho^b \rho^c$ and $f_{\alpha\beta c}\lambda^{\alpha} \lambda^{\beta} \rho^c$
immediately
come to mind. Defining
$$
V={1\over (C_A)^2}\,\,\Bigl({1\over 3}f_{abc}\rho^a \rho^b \rho^c +
f_{\alpha\beta c}\lambda^{\alpha} \lambda^{\beta} \rho^c\Bigr),
\eqn\zai
$$
one can check that
$$
Q(z)V(w)={1\over z-w}W(w),
\eqn\zaj
$$
which proves our point: $(V,W)$ forms a new BRST doublet.
In trying to work out the topological
structure present in the theory, we have
found that the operator algebra is very similar
to that of the twisted $N=2$ models.
But the appearance of $(V,W)$ forces us to compute
their OPE's with all other
operators, and there is no guarantee that
the resulting algebra will close on
a finite number of fields. However, the
algebra of $(G,T)$, $(R,Q)$ and
$(V,W)$ does close, as some computation proves. The results are
$$
\eqalign{
T(z)W(w)=&{3\over (z-w)^2}W(w) + {1\over z-w}\partial W(w)\cr
T(z)V(w)=&{3\over (z-w)^2}V(w) + {1\over z-w}\partial V(w)\cr
G(z)W(w)=&{3\over (z-w)^2}V(w) + {1\over z-w}\partial V(w)\cr
R(z)W(w)=&{-2\over z-w}W(w)\cr
R(z)V(w)=&{-3\over z-w}V(w),\cr}
\eqn\zak
$$
while all other OPE's vanish identically.
It is important to emphasise that,
contrary to what happens with ${\cal W}$ algebras,
the existence of
higher-spin fields does not spoil the linearity of the algebra.
The above conclusions have also been obtained in
[\tca,\gln], but our
analysis here extends these results to the
more general case of an arbitrary
superalgebra. It should also be mentioned
that the algebra exhibited in
eqs. \zaa\ to \zak, which we shall call the Kazama
algebra, first appeared in [\kazama] as a consistent,
non-trivial extension of the twisted $N=2$ algebra.
In [\kazama] it was related to an $N=1$
superconformal symmetry, but no explicit
representation for the generators was given
(see also ref.
\REF\getzler{E. Getzler
\journalAnn. Phys.&237(95)161.}[\getzler]).
To complete our analysis, it remains to
study whether or not the currents
${\cal J}_a$, ${\cal J}_{\alpha}$ and their
BRST ancestors $\rho_a$, $\lambda_{\alpha}$,
on the one hand, and the generators $T$,
$G$, $R$, $Q$, $W$ and $V$,
on the other, are compatible. Some of the
corresponding OPE's are trivial
(for example, those expressing the Virasoro primary
character of the currents); others have already
been given (eq. \zw).
Among those remaining, the only non-vanishing ones are
$$
\eqalign{
G(z){\cal J}_a(w)=&{1\over (z-w)^2}\rho_a(w) +
{1\over z-w}\partial \rho_a(w)\cr
G(z){\cal J}_{\alpha}(w)=&{1\over (z-w)^2}\lambda_{\alpha}(w) +
{1\over z-w}\partial \lambda_{\alpha}(w)\cr
R(z)\rho_a(w)=&{-1\over z-w} \rho_a(w)\cr
R(z)\lambda_{\alpha}(w)=&{-1\over z-w}\lambda_{\alpha}(w),\cr}
\eqn\zal
$$
which establishes that the topological and
current superalgebra structures are
indeed compatible.
An interesting feature of the above
construction is the fact that it can also
be performed with two independent sets of
currents $J^{1}$, $J^{2}$. Suppose
$J^{1}$ and $J^{2}$ satisfy the algebra \ze\
with levels $k_1$ and $k_2$,
respectively. Then the Sugawara energy-momentum
tensor $T^J$ is given by
$$
T^J=
{1\over 2k_1 + C_A}(g^{ab}J^{1}_{a}J^{1}_{b} +
g^{\alpha\beta}J^{1}_{\alpha}J^{1}_{\beta}) +
{1\over 2k_2 + C_A}(g^{ab}J^{2}_{a}J^{2}_{b} +
g^{\alpha\beta}J^{2}_{\alpha}J^{2}_{\beta}),
\eqn\zam
$$
and the corresponding central charge is
$$
c_J= [\,{2k_1 \over 2k_1+ C_A}\,
+\, {2k_2 \over 2k_2+ C_A }]\,(d_B - d_F).
\eqn\zan
$$
Imposing $k_1+k_2=-C_A$ and
setting $k_1=k$ for simplicity,
eqs. \zam\ and \zan\ reduce to
$$
T^J=
{1\over 2k + C_A}\bigr[g^{ab}(J^{1}_{a}J^{1}_{b}-
J^{2}_{a}J^{2}_{b})+
g^{\alpha\beta}(J^{1}_{\alpha}J^{1}_{\beta} -J^{2}_{\alpha}J^{2}_{\beta})\bigl]
\eqn\zao
$$
and
$$
c_J=2(d_B-d_F),
\eqn\zap
$$
which exactly cancels the ghost central
charge $c_{(\rho\gamma)}+c_{(\eta\lambda)}$. That
this is indeed a new topological point
is again confirmed by the following
arguments. The new total currents are
$$
\eqalign{
{\cal J}_a=&J^1_a + J^2_a + J^{{\rm gh}}_a\cr
{\cal J}_{\alpha}=&J^1_{\alpha} + J^2_{\alpha} + J^{{\rm gh}}_{\alpha},\cr}
\eqn\zaq
$$
with $J^{{\rm gh}}$ as in eqs. \zr, \zs, and their algebra
is $$
\eqalign{
{\cal J}_a(z){\cal J}_b(w)=&{(k_1+k_2+C_A) g_{ab}
\over (z-w)^2} +
{f_{ab}^{c}\over z-w}{\cal J}_c(w)\cr
{\cal J}_a(z){\cal J}_{\beta}=&{f_{a\beta}^{\gamma}\over z-w}
{\cal J}_{\gamma}(w)\cr
{\cal J}_{\alpha}(z){\cal J}_{\beta}(w)=&{(k_1+k_2+C_A)
g_{\alpha\beta}\over (z-w)^2} +
{f_{\alpha\beta}^{c}\over z-w}{\cal J}_c(w).\cr}
\eqn\zar
$$
The new BRST current making them BRST-exact is
$$
Q=-\gamma^a (J^1_a+J^2_a) - {1\over 2}f_{ab}^{c}\gamma^a\gamma^b\rho_c +
f_{a\beta}^{\mu}\gamma^a\eta^{\beta}\lambda_{\mu} - \eta^{\alpha} (J^1_{\alpha}+J^2_{\alpha}) -
{1\over 2}f_{\alpha\beta}^{c}\eta^{\alpha}\eta^{\beta}\rho_c,
\eqn\zas
$$
where again $\rho_a$ and $\lambda_{\alpha}$ are their BRST ancestors.
Nilpotency of $Q$
occurs only at the topological point, since
$$
Q(z)Q(w) = {k_1+k_2+C_A\over z-w}\,(g_{ab}\partial\gamma^a\gamma^b +
g_{\alpha\beta}\partial \eta^{\alpha}\eta^{\beta}).
\eqn\zat
$$
One can now repeat the above analysis and work out the expressions for all the
operators, with the result that the
topological algebra is satisfied without
changes. The ghost number current $R$ and the
topological dimension $d$ remain the same, but the
other generators have to be modified as follows:
$$
\eqalign{
G=& {1\over 2k+C_A}\,\, \Bigl[g^{ab}\rho_a (J^1_b-J^2_b) +
g^{\alpha\beta}\lambda_{\alpha}(J^1_{\beta}-J^2_{\beta})\Bigr]\cr
W=&{1\over (2k+C_A)^2}\,\,\Bigl[-C_A\,(\partial \rho_a\rho^a +
\partial \lambda_{\alpha}\lambda^{\alpha}) +
2 f_{a\beta}^{\gamma} \rho^a \lambda^{\beta} (J^1_{\gamma}+J^2_{\gamma})\cr
+&f_{ab}^c\rho^a \rho^b (J^1_c+J^2_c) +
f_{\alpha \beta}^c \lambda^{\alpha} \lambda^{\beta} (J^1_c+J^2_c)\Bigr]
\cr
V=&{1\over (2k+C_A)^2}\,\,
\Bigl[{1\over 3}f_{abc}\rho^a \rho^b \rho^c +
f_{\alpha\beta c}\lambda^{\alpha} \lambda^{\beta} \rho^c\Bigr].\cr}
\eqn\zau
$$
Although the topological algebra is
the same as in the one-current case, this
two-current construction is interesting
because it allows for a lagrangian
interpretation of the theory as a
${\cal G}/{\cal G}$ coset. This point is
examined in the next section.
Before finishing this one, let us point out that the topological
algebra we have studied admits deformations both in its one and two
current realizations. Indeed, if $\alpha^a$ are c-number constants, one
can redefine $T$, $G$ and $R$ as follows:
$$
\eqalign{
T&\rightarrow T+\sum_a\alpha^a\partial {\cal J}_a\cr
G&\rightarrow G+\sum_a\alpha^a\partial \rho_a\cr
R&\rightarrow R+\sum_a\alpha^a {\cal J}_a,\cr}
\eqn\zauu
$$
The operators $Q$, $V$ and $W$ are left unaffected by the deformation.
One easily checks that the transformed generators satisfy the extended
topological algebra for any value of the $\alpha^a$ constants. Of
course, after the deformation, the currents are no longer primary
dimension-one operators. Transformations of the type displayed in
eq.\zauu\ will play an important role in section 4, where we shall
relate our results with the non-critical string theories.
\chapter{A gauged, supergroup-valued WZW model.}
The topological algebra described in
the previous section has a lagrangian interpretation that
we now develop. We shall show below that it is possible to give a
lagrangian description of the two-current
construction of section 2. The main
result of this section is the
interpretation of the gauged, ${\cal G}$-valued WZW model as a
theory in which the extended topological algebra closed by
$(T,G)$, $(Q,R)$ and $(W,V)$ is realised. A similar conclusion has
also been reported in [\tca] for the bosonic case (\ie, when
$d_F=0$), but our presentation here is totally general. An earlier
reference on gauged WZW models is \REF\schnitzer{H. J.
Schnitzer\journalNucl. Phys.&B324(89)412.} [\schnitzer]. For arbitrary
supergroups, a lagrangian construction of ${\cal G}/{\cal G}$ has
already been put forward \REF\yu{J. B. Fan and M. Yu, ``G/G Gauged
Supergroup Valued WZNW Field Theory", Academia Sinica preprint
AS-ITP-93-22.} in [\yu].
Our starting point is the ${\cal G}$-gauged WZW functional
$$
\Gamma (g,A) = \,\Gamma (g) -
{1\over \pi}\int_{\Sigma}d^2z\,
{\rm str}(g^{-1}A_{\bar z}gA_z - A_{\bar z}\partial_zg g^{-1} +
g^{-1}\partial_{\bar z}gA_z - A_zA_{\bar z}),
\eqn\zav
$$
with $\Gamma (g)$ given by
$$
\Gamma (g) = {1\over 2 \pi}\int_{\Sigma}d^2z\,{\rm
str}(g^{-1}\partial_zgg^{-1}\partial_{\bar z}g) + {i\over
12\pi}\int_{M}\epsilon^{\mu\nu\rho}\,{\rm
str}(g^{-1}\partial_{\mu}g
g^{-1}\partial_{\nu}gg^{-1}\partial_{\rho}g).
\eqn\zaw
$$
$g$ is a function taking values in
the supergroup whose Lie superalgebra is ${\cal G}$, and the
3-manifold $M$ is such that $\partial M = \Sigma$.
The {\it na\"{\i}ve} partition function is
$$
Z = \int Dg DA_{\bar z}DA_z \exp{[-k \Gamma (g,A)]}.
\eqn\zax
$$
The gauge invariance of $\Gamma (g,A)$ is well known.
Also useful is the Polyakov-Wiegmann identity
satisfied by $\Gamma (g)$,
$$
\Gamma (gh) = \Gamma (g) + \Gamma (h) + \langle g,h\rangle ,
\eqn\zay
$$
where
$$
\langle g,h\rangle = {1\over \pi}\int_{\Sigma}d^2z\,
{\rm str}(g^{-1}\partial_{\bar z}g\,\partial_z
hh^{-1}).
\eqn\zaz
$$
Parametrise the gauge fields as
$$
A_{\bar z}=h^{-1}\partial_{\bar z}h\qquad
A_z={\bar h}^{-1}\partial_z \bar h
\eqn\zba
$$
with $h$ and $\bar h$ taking values in the supergroup,
and change variables in the functional
integral \zax. One has
$$
DA_zDA_{\bar z}=\,J[h,\bar h]\,DhD\bar h,
\eqn\zbb
$$
where $J[h,\bar h]$ is the Jacobian for the change of
variables $A_z,\,\, A_{\bar z}\rightarrow h,\,\, \bar h$.
This Jacobian can be represented as a functional integral
over ghost fields that take values in the adjoint
representation of ${\cal G}$. Denoting, as in the
previous section, these ghost fields by $(\rho_a,\gamma^a)$
and $(\lambda_{\alpha},\eta^{\alpha})$,
we have:
$$
J[h,\bar h]\,=\,\exp {[C_A
\Gamma (h^{-1}\bar h)]}\,
\int D\rho D\gamma D\lambda D\eta \,
\exp{[{-1\over \pi}\int_{\Sigma} d^2z\,
(\rho_a\partial_{\bar z}\gamma^a +
\lambda_{\alpha}\partial_{\bar z}\eta ^{\alpha} + {\rm
c.c.})]}.
\eqn\extra
$$
Taking into account that
$$
\Gamma (g,A) = \,\Gamma (h^{-1}g\bar h) -
\Gamma (h^{-1}\bar h),
\eqn\extrados
$$
the partition function becomes
$$
\eqalign{
Z=\int DgDhD\bar h D\rho D\gamma D\lambda D\eta
\,&\exp{[-k \Gamma (h^{-1}g\bar h) +
(k+ C_A)\Gamma(h^{-1}\bar h)]}\cr
&\exp{[{-1\over \pi}\int_{\Sigma} d^2z\,
(\rho_a\partial_{\bar z}\gamma^a +
\lambda_{\alpha}\partial_{\bar z}\eta ^{\alpha} +
{\rm c.c.})]}.\cr}
\eqn\zbc
$$
Changing variables as $h^{-1}g\bar h\rightarrow g$ and
choosing the gauge $h=1$ (which does not introduce any
new Faddeev-Popov ghosts) we obtain:
$$
\eqalign{
Z=\int DgD\bar h D\rho D\gamma D\lambda D\eta
&\exp{[-k \Gamma (g) + (k+C_A)\Gamma(\bar h)]}\cr
&\exp{[{-1\over \pi}\int_{\Sigma} d^2z\,
(\rho_a\partial_{\bar z}\gamma^a +
\lambda_{\alpha}\partial_{\bar z}\eta ^{\alpha} +
{\rm c.c.})]}.\cr}
\eqn\zbd
$$
This gauge-fixed form for the partition function clearly
exhibits the necessary elements to
construct the ${\cal G}/{\cal G}$ theory,
as realised with two currents: one needs two independent,
ungauged WZW models such that their levels add up to $-C_A$, plus a
compensating ghost sector in order to set the total level to zero. Let
us finally point out that the lagrangian interpretation we have
discussed in this section allows to interpret the BRST symmetry of the
superalgebra ${\cal G}$ as the basic symmetry of a topological sigma
model having a Lie supergroup as target space.
\chapter{Relation with non-critical superstrings.}
In this section we shall explore the relation between the
topological theories constructed in the previous sections and the
non-critical superstring theories. In particular, we will argue
that, for the Lie superalgebras ${\rm osp}(1|2)$ and ${\rm osp}(2|2)$, the
corresponding topological coset models are related to the $N=1$
and $N=2$ superstring theories, respectively. In order to find
this correspondence one must first conveniently deform the
two-current model, as was explained at the end of section 2.
In the deformed theory one can implement a quantum
Drinfeld-Sokolov hamiltonian reduction in such a way that the
reduced model can be identified with the corresponding string
theory.
The connection of the topological current system with non-critical
strings is more transparent if a free field realisation of the
two currents involved in the ${\cal G}/ {\cal G}$ coset is
used. Roughly speaking, one can associate one of the two currents
with the matter sector of the string whereas the other current is
related to the Liouville degrees of freedom. Moreover,
some of the ghosts of the deformed ${\cal G}/ {\cal
G}$ coset can be identified with those of string theory. The
remaining fields coming from the ghost and current sectors can be
organised into topological quartets and one can invoke the
standard Kugo-Ojima confinement mechanism to eliminate
these quartets from the physical Hilbert space.
Before studying the ${\rm osp}(1|2)$ and ${\rm osp}(2|2)$ cases, let us
for completeness
recall
[\yank,\hu] the relation between the ${\rm sl}(N)/{\rm sl}(N)$
cosets and the non-critical $W_N$-strings. First of all let us
briefly describe the root system of the ${\rm sl}(N)$ Lie algebra. If
$\vec e_i$ is a unitary vector in ${{\rm I\kern-1.6pt {\rm R}}}^N$ along the ${\rm
i}^{{\rm th}}$ axis and $\vec \epsilon_{ij}\,=\,\vec e_i\,-\,\vec
e_j$, then the positive roots of ${\rm sl}(N)$ are the elements of the set
$\Delta_+\,=\,\{\,\vec \epsilon_{ij},\,\,\,j>i\,\}$, while the
simple roots are given by
$\vec \alpha_i\,=\,\vec\epsilon_{i,i+1}$ ($i=1,\cdots,N-1)$. Any
positive root $\vec \alpha\in\Delta_+$ can be written as
$\vec \alpha\,=\,\sum_{i=1}^{N-1}\,n_{\alpha}^i\vec\alpha_i$
where the $n_{\alpha}^i$'s are non-negative integers. The height
of $\vec \alpha$ is given by:
$$
h_{\alpha}\,=\,\sum_{i=1}^{N-1}\,n_{\alpha}^i.
\eqn\uno
$$
In particular, for $j>i$, $h_{\epsilon_{ij}}\,=\,j-i$. Denoting
by $\vec \delta$ half the sum of positive roots
(\ie\
$\vec \delta\,=\,{1\over 2}\,\sum_{\alpha\in\Delta_+}\,
\vec \alpha$), the height of any $\vec \alpha\in\Delta_+$ is
simply given by
$h_{\alpha}\,=\,\vec\delta\cdot\vec\alpha$.
We adopt the free field realisation of the ${\rm sl}(N)$ current
algebra of ref.
\REF\gera{A. Gerasimov et al. \journalInt. J. Mod. Phys.&A5(90)2495.}
[\gera]. In this realisation, a spin-one
bosonic $\beta\gamma$ system is introduced for every positive
root $\vec \alpha\in\Delta_+$. If we call
these bosonic fields
$(w_{\alpha},
\chi_{\alpha})$, then the expression for the
currents associated with the positive roots is of the form:
$$
J_{\alpha}\,=\,w_{\alpha}+\cdots,
\eqn\dos
$$
where the dots denote terms which are non-linear in the fields.
Notice that the conformal weights of $(w_{\alpha},
\chi_{\alpha})$ are $\Delta (w_{\alpha})=1$ and
$\Delta (\chi_{\alpha})=0$. One also needs to introduce a set of
$N-1$
scalar fields $\vec \phi\,=\,(\phi_1,\cdots, \phi_{N-1}\,)$.
Then the expression for the Cartan currents can be
given in general. In fact if we represent by $\vec\mu\cdot\vec H$
the current along the direction of an arbitrary Cartan vector
$\vec\mu$, we have:
$$
\vec H\,=\,i\sqrt{k+N}\,\,\partial\vec\phi\,-\,
\sum_{\alpha\in\Delta_+}\,\vec\alpha\,w_{\alpha}
\chi_{\alpha},
\eqn\tres
$$
where $k$ is the level of the ${\rm sl}(N)$ algebra. It is also
possible to give a simple expression for the Sugawara
energy-momentum tensor in terms of these free fields:
$$
\eqalign{
T^J\,=&\,{1\over
2(k+N)}\,g^{ab}J_{a}J_{b}=\cr
=&\sum_{\alpha\in\Delta_+}\,w_{\alpha}\partial\chi_{\alpha}
-{1\over 2}\,(\partial \vec\phi)^2\,-\,
{i\over\sqrt{k+N}}
\,\vec\delta\cdot\partial^2\vec\phi.\cr}
\eqn\cuatro
$$
A simple calculation using the Freudenthal-de Vries ``strange"
formula ($12\, \vec \delta^2=N(N^2-1)$) shows that
the energy-momentum tensor in eq. \cuatro\ indeed has the correct
central charge for an affine algebra at level $k$.
The topological ${\rm sl}(N)/ {\rm sl}(N)$ coset is obtained by
combining two ${\rm sl}(N)$ currents with levels $k_1=k$ and
$k_2=-k-2N$. Let the corresponding free fields carry the labels
$1$ and $2$. We must also add a pair of fermionic ghosts for each
independent current direction. In what follows we shall denote
the ghost along the Cartan direction by $(\rho_i, \gamma^i)$
($i=1,\cdots,N-1$) and those associated with the positive
(negative) roots of the algebra by
$(\rho_{\alpha},\gamma^{\alpha})$
( $(\rho_{-\alpha},\gamma^{-\alpha})$ respectively).
In order to make contact with non-critical string theory we must
first deform the theory. It turns out that the appropriate
deformation of the total energy-momentum tensor $T$ is:
$$
T_{{\rm improved}}\,=\,T\,+\,\vec\delta\cdot\partial\vec
{\cal H}.
\eqn\cinco
$$
In eq. \cinco\ $\vec {\cal H}$ is the total Cartan current of the
${\rm sl}(N)/ {\rm sl}(N)$ coset (see eq. \zaq).
Let us separate in $T_{{\rm improved}}$ the contribution of the
currents from those of the ghosts:
$$
T_{{\rm improved}}\,=\,T_{{\rm improved}}^J+
T_{{\rm improved}}^{\rm gh}.
\eqn\seis
$$
Using the free-field representation of $T^J$ and
$\vec H$ given in eqs. \cuatro\ and \tres\ we can write the
explicit expression of $T_{{\rm improved}}^J$:
$$
\eqalign{
T_{{\rm improved}}^J\,=&\,
-{1\over 2}\,( \partial \vec\phi_1)^2\,
-{1\over 2}\,( \partial \vec\phi_2)^2\,+
i{t-1\over \sqrt{t}}\,\vec\delta\cdot\partial^2\vec\phi_1
\,-\,{t+1\over\sqrt{t}}\,
\vec\delta\cdot\partial^2\vec\phi_2\,+\cr
+&\,\sum_{i=1,2}\,\,\sum_{\alpha\in\Delta_+}[\,
(1-h_{\alpha})\,w_{\alpha}^i\partial\chi_{\alpha}^i
\,-\,h_{\alpha}\,\partial
w_{\alpha}^i\chi_{\alpha}^i\,],\cr}
\eqn\siete
$$
where $t=k+N$. Notice that, in the deformed theory, the fields
$(w_{\alpha},\chi_{\alpha})$ acquire a conformal weight that
depends on the height of the root $\vec \alpha$
($\Delta (w_{\alpha})=1-h_{\alpha}$,
$\Delta (\chi_{\alpha})=h_{\alpha}$). In order to compute the
ghost contribution to the improved energy-momentum tensor, we
need to know the part of $\vec {\cal H}$ that depends on the ghost
fields. From the commutation relations of ${\rm sl}(N)$ one easily gets:
$$
\vec {\cal H}^{\rm gh}\,=
\,\sum_{\alpha\in\Delta_+}\,\vec \alpha\,
[\,\gamma^{\alpha}\rho_{\alpha}-
\gamma^{-\alpha}\rho_{-\alpha}\,].
\eqn\ocho
$$
Using eq. \ocho\ one obtains
$$
\eqalign{
T_{{\rm improved}}^{\rm gh}=&\,\sum_{i=1}^{N-1}
\rho_i\partial \gamma^i\,+\,
\sum_{\alpha\in\Delta_+}[\,(1-h_{\alpha})\,\rho_{\alpha}
\partial\gamma^{\alpha}\,+\,h_{\alpha}\gamma^{\alpha}
\partial\rho_{\alpha}\,]\,+\cr
+&\sum_{\alpha\in\Delta_+}[\,(1+h_{\alpha})\,\rho_{-\alpha}
\partial\gamma^{-\alpha}\,-\,h_{\alpha}\gamma^{-\alpha}
\partial\rho_{-\alpha}\,].\,\cr}
\eqn\nueve
$$
The central charge of the field $\vec\phi_1$ can be computed from
the background charge displayed in eq. \siete. A simple
calculation shows that
$$
c_{\phi_1}\,=\,(N-1)\,[1\,-\,N(N+1)\,{(t-1)^2\over t}\,].
\eqn\diez
$$
When $t\,=\,{q\over p}$ with $q,p\in {\rm Z}\kern-3.8pt {\rm Z} \kern2pt$ (\ie\ when
$k+N={q\over p}$), the central charge in eq. \diez\ is precisely
that of the minimal $(p,q)$ model of $W_N$ matter. It is also
easy to check from eq. \diez\ that $\vec \phi_2$ has
the correct background charge to be considered as the
$W_N$-Liouville field. Moreover it can be seen that one can
always combine in a Kugo-Ojima topological quartet the $(w_{\alpha}^i,
\chi_{\alpha}^i)$ systems with ghost fields having the same conformal
weights. We can pair, for example, the $(w_{\alpha}^1, \chi_{\alpha}^1)$
fields with the ghosts $(\rho_{\alpha}, \gamma^{\alpha})$ corresponding
to the positive roots. Also the Cartan ghosts
$(\rho_i, \gamma^i)$ can be paired with the fields
$(w_{\alpha}^2, \chi_{\alpha}^2)$ when $\vec\alpha$ is a simple
root (\ie\ when $h_{\alpha}=1$), since in this case both systems
have dimensions $(1,0)$ and there are equal number of them. The
$(w_{\alpha}^2, \chi_{\alpha}^2)$ fields with $h_{\alpha}\geq 2$ can be
paired with some of the $(\rho_{-\alpha}, \gamma^{-\alpha})$ ghosts.
By looking at the conformal weights of these last two systems one
concludes that, in order to combine $(w_{\alpha}^2, \chi_{\alpha}^2)$
with $(\rho_{-\alpha'}, \gamma^{-\alpha'})$, the heights of $\vec
\alpha$ and $\vec\alpha'$ must satisfy
$h_{\alpha}-h_{\alpha'}=1$. A simple calculation tells one
how many fields can be paired in this way. Since
the number of roots of ${\rm sl}(N)$ with height $h$ is $N-h$, the
difference between the number of
$(\rho_{-\alpha'}, \gamma^{-\alpha'})$ and
$(w_{\alpha}^2, \chi_{\alpha}^2)$ systems is
$N-h_{\alpha'}-(N-h_{\alpha})=h_{\alpha}-h_{\alpha'}=1$. It
follows that, for a given height $h$, there always remains one
unpaired $(\rho,\gamma)$ system with conformal weights
$(1+h,-h)$. We are thus left with a set of $N-1$
anticommuting ghost fields
with conformal weights $(2,-1),\cdots, (N,1-N)$.
Let us denote these fields by $(b_j,c_j)$ where the
conformal weight of $b_j$ is $j+1$ for $j=1,\cdots,N-1$. Notice
that they correspond to the ghost system of the $W_N$ string.
Therefore, writing $\vec \phi_M$ ($\vec \phi_L$) instead
of $\vec\phi_1$ ($\vec\phi_2$ respectively),
the reduced energy-momentum tensor is given by:
$$
\eqalign{
T_{{\rm reduced}}\,=&\,
-{1\over 2}\,( \partial \vec\phi_M)^2\,
-{1\over 2}\,( \partial \vec\phi_L)^2\,+
i{t-1\over \sqrt{t}}\,\vec\delta\cdot\partial^2\vec\phi_M
\,-\,{t+1\over\sqrt{t}}\,
\vec\delta\cdot\partial^2\vec\phi_L\,+\cr
+&\,\sum_{j=1}^{N-1}\,\,[\,(j+1)b_j\partial c_j\,-\,jc_j\partial
b_j\,],\cr}
\eqn\once
$$
which, as stated above, corresponds to that
of matter coupled to $W_N$-gravity.
Next let us consider the ${\rm osp}(1|2)$ Lie superalgebra. This
algebra contains three bosonic currents $J_{\pm}$ and $H$ that
close an ${\rm sl}(2)$ algebra at level $k$. In addition there are
two fermionic currents that we shall denote by
$j_{\pm}$. The affine ${\rm osp}(1|2)$ superalgebra can be realised
[\bershadsky]
in terms of one scalar field $\phi$, one bosonic $\beta\gamma$ system
(denoted by $(w,\chi)$), and one fermionic $bc$ system
(denoted by $(\bar\psi, \psi)$), with dimensoins
$\Delta(w)=\Delta (\bar\psi)=1$ and
$\Delta(\chi)=\Delta (\psi)=0$, satisfying the following basic OPE's:
$$
w(z)\,\chi(w)\,=\,\psi(z)\,\bar\psi(w)\,=\,{1\over z-w}
\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
\phi(z)\,\phi(w)\,=\,-{\rm log}\,(z-w).
\eqn\doce
$$
In terms of these fields the explicit form of the ${\rm osp}(1|2)$
currents is:
$$
\eqalign{
J_+\,=&\,w\cr
J_-\,=&-\,w\chi^2\,+\,i\sqrt{2k+3}\,\,\chi\partial\phi\,-
\,\chi\psi\bar\psi\,+k\partial\chi\,+
\,(k+1)\psi\partial\psi\cr
H\,=&-w\chi\,+{i\over 2}\,\sqrt{2k+3}\,\,\partial\phi\,-\,
{1\over 2}\,\psi\bar\psi\cr
j_+\,=&\bar\psi\,+\,w\psi\cr
j_-\,=&-\chi(\bar\psi\,+\,w\psi)\,+i\sqrt{2k+3}\,\,
\psi\partial\phi\,+\,(2k+1)\partial\psi.\cr}
\eqn\trece
$$
Notice that for ${\rm osp}(1|2)$, with our
conventions, $C_A=3$. Using this
value and the metric tensor extracted from OPE's of the currents in eq.
\trece, we can write the Sugawara energy-momentum tensor:
$$
T^J\,=\,{1\over
2k+3}\,[\,J_+J_-\,+\,J_-J_+\,+\, 2H^2\,-\,{1\over
2}\,j_+\,j_-\,+{1\over 2}\,j_-j_+\,].
\eqn\catorce
$$
Substituting the representation given in eq. \trece\ into
eq.\catorce\ one gets:
$$
T^J\,=\,w\partial\chi\,-\,\bar\psi\partial
\psi\,-\,{1\over 2}\,(\partial\phi)^2\,-\,
{i\over 2\sqrt{2k+3}}\,\partial^2\phi.
\eqn\quince
$$
In order to realise the topological ${\rm osp}(1|2)/{\rm osp}(1|2)$ coset
model we must combine two current systems with levels $k_1=k$
and $k_2=-k-3$. Let us denote the anticommuting ghost systems
for the currents $H$ and $J_{\pm}$ by $(\rho_0,\gamma^0)$ and
$(\rho_{\pm},\gamma^{\pm})$ respectively. The commuting ghosts
associated to the $j_{\pm}$ currents will be similarly denoted by
$(\lambda_{\pm},\eta^{\pm})$. In complete parallel with the
${\rm sl}(N)$ case, let us deform the energy-momentum tensor by adding
a derivative along the total Cartan current ${\cal H}$:
$$
T_{{\rm improved}}\,=\,T_{{\rm improved}}^{\rm J}\,+\,
T_{{\rm improved}}^{\rm gh}\,\equiv\,T\,
+\,\partial{\cal H}.
\eqn\dseis
$$
In eq. \dseis\ we have separated the contributions of the
currents from those of the ghosts. Let us consider
$T_{{\rm improved}}^{\rm J}$ first. Using eq. \quince\ one
immediately arrives at:
$$
\eqalign{
T_{{\rm improved}}^{\rm J}\,=&\,
-{1\over 2}\,(\partial\phi_1)^2\,
-\,{1\over 2}\,(\partial\phi_2)^2\,+\,
{i\over 2}\,{t-1\over\sqrt t}\,\partial^2\phi_1\,-\,
{1\over 2}\,{t+1\over\sqrt t}\,\partial^2\phi_2\,-\cr
-&\sum_{i=1,2}\,[\partial w^i\chi^i\,+\, {1\over 2}
\bar \psi^i\partial\psi^i\
+\, {1\over 2}
\psi^i\partial\bar\psi^i\,],\cr}
\eqn\dsiete
$$
where now $t\,=\,2k+3$ and, as in the ${\rm sl}(N)$ case,
the indices $1$ and $2$ label the two currents. On the other
hand, from the basic OPE's of ${\rm osp}(1|2)$, the contribution of the
ghost fields to ${\cal H}$ is easily obtained. One gets:
$$
{\cal H}\,^{{\rm gh}}=
\,\gamma^+\rho_+\,-\,\gamma^-\rho_-\,- {1\over
2}\,\eta^+\lambda_+\,+\, {1\over 2}\,\eta^-\lambda_-.
\eqn\docho
$$
Taking eq. \docho\ into account it is straightforward to compute
$T_{{\rm improved}}^{\rm gh}\,$:
$$
\eqalign{
T_{{\rm improved}}^{{\rm gh}}\,=&
\,\rho_0\partial\gamma^0\,+\,
\gamma^+\partial\rho_+\,+\,2\rho_-\partial\gamma^- \,-\,
\gamma^-\partial\rho_-\cr
+&\,{1\over 2}\partial \eta^+\lambda_+
-\,{1\over 2} \eta^+\partial\lambda_+\,+\,
\,{3\over 2}\partial \eta^-\lambda_-\,+\,{1\over 2}
\eta^-\partial\lambda_-.\cr}
\eqn\dnueve
$$
Notice that in the deformed theory the conformal weights of the
antighost fields are:
$$
\Delta (\rho_0)\,=\,1
\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\rho_+)\,=\,0
\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\rho_-)\,=\,2
\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\lambda_+)\,=\,{1\over 2}
\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\lambda_-)\,=\,{3\over 2}.
\eqn\veinte
$$
Therefore $(\rho_-,\gamma^-)$ and $(\lambda_-,\eta^-)$ have
acquired the right dimensions to become the superdiffeomorphism
ghosts of the $N=1$ string. A glance at eqs. \dnueve\ and
\dsiete\ reveals that the other ghosts can be accommodated in
quartets with fields coming from the current sector. Indeed one
can pair $(\rho_0,\gamma^0)$ and $(\rho_+,\gamma^+)$ with
$(w^1,\chi^1)$ and $(w^2,\chi^2)$. The commuting ghosts
$(\lambda_+,\eta^+)$ can be paired with the
$({1\over 2},{1\over 2})$ fermionic system obtained from, say, the
fields ${1\over \sqrt{2}}\,(\psi^1+\bar\psi^1)$ and
${1\over \sqrt{2}}\,(\psi^2+\bar\psi^2)$.
After this process there remain two Majorana fields
$\psi_M={i\over \sqrt{2}}\,(\psi^1-\bar\psi^1)$ and
$\psi_L={i\over \sqrt{2}}\,(\psi^2-\bar\psi^2)$. Calling $\phi_M$,
$\phi_L$, $b$, $c$, $\beta$ and $\gamma$
to $\phi_1$, $\phi_2$, $\rho_-$, $\gamma^-$, $\lambda_-$ and
$\eta^-$ respectively, we can write
the reduced energy-momentum tensor as:
$$
\eqalign{
T_{{\rm reduced}}\,=&\,
-{1\over 2}\,(\partial\phi_M)^2\,
-\,{1\over 2}\,(\partial\phi_L)^2\,+\,
{i\over 2}\,{t-1\over\sqrt t}\,\partial^2\phi_M\,-\,
{1\over 2}\,{t+1\over\sqrt t}\,\partial^2\phi_L\,-\cr
&-{1\over 2}\,\psi_M\partial\psi_M\,-\,
{1\over 2}\,\psi_L\partial\psi_L\,
+\,2b\partial c\,-\,c\partial b\,+\,{3\over 2}\partial\gamma
\beta\,+\,{1\over 2}\,\gamma\partial\beta,\cr}
\eqn\vuno
$$
which is indeed the one corresponding to the $N=1$ RNS
superstring. Furthermore the matter central charge in
$T_{{\rm reduced}}$ is:
$$
c_M\,=\,{3\over2}\,(\,1\,-\,2\,{(t-1)^2\over t}\,).
\eqn\vdos
$$
When $t=2k+3={q\over p}$ with $p,q\in {\rm Z}\kern-3.8pt {\rm Z} \kern2pt$, eq. \vdos\ gives the
central charge of the minimal models of the $N=1$ superconformal
symmetry.
To finish this section let us now analyse the ${\rm
osp}(2|2)$ current algebra. This algebra contains four
bosonic currents ($J_{\pm}$, $H$ and $J$) along with
other four fermionic ones ($j_{\pm\pm}$). The currents
$J_{\pm}$ and $H$ close an ${\rm sl}(2)$ algebra, while
$J$ is a $U(1)$ current. One can represent this algebra by
means of two scalar fields $\phi$ and $\varphi$, one
$(1,0)$ commuting $\beta\gamma$ system (denoted by
$(w,\chi)$) and two $(1,0)$ anticommuting $bc$ systems
($(\bar\psi_+,\psi_-)$ and $(\bar\psi_-,\psi_+)$). We
shall use the conventions of eq. \doce\ for the OPE's of
the fields $(w,\chi)$, $\phi$ and $\varphi$. For the
fermionic fields, the basic OPE's are:
$$
\psi_+(z)\bar\psi_{-}(w)\,=\,\psi_-(z)\bar\psi_+(w)\,=\,
{1\over z-w}.
\eqn\vtres
$$
Then the explicit representation[\bershadsky] of the ${\rm osp}(2|2)$
currents is:
$$
\eqalign{
J_+\,=&\,w\cr
J_-\,=&-w\chi^2\,+\,i\sqrt{2k+2}\,\,\chi\partial\phi\,
-\chi(\,\psi_-\bar\psi_+\,+\,\psi_+\bar\psi_-\,)\cr
-&\sqrt{2k+2}\,\,\psi_+\psi_-\partial\varphi\,+\,
k\partial\chi\,+\,(k+1)[\,\psi_-\partial\psi_+\,+\,
\,\psi_+\partial\psi_-\,]\cr
H\,=&-w\chi\,+\,{i\over 2}\,\sqrt{2k+2}\,\,\partial\phi\,
-\,{1\over 2}\,
[\psi_-\bar\psi_+\,+\,\psi_+\bar\psi_-\,]\cr
J\,=&\,-\,{1\over 2}\,
[\psi_-\bar\psi_+\,-\,\psi_+\bar\psi_-\,]
\,+\,{\sqrt{2k+2}\over 2}\,\,\partial\varphi\cr
j_{+\pm}\,=&\,\bar\psi_{\pm}\,+\,w\psi_{\pm}\cr
j_{-\pm}\,=&\,-\chi\,(\bar\psi_{\pm}\,+\,w\psi_{\pm}\,)
\,+\,\sqrt{2k+2}\,\,\psi_{\pm}\, \partial(i\phi\,\pm\,
\varphi\,)\,+\,(2k+1)\,\partial\psi_{\pm}\,+\,
\psi_{\pm}\psi_{\mp}\bar\psi_{\pm}.\cr\cr}
\eqn\vcuatro
$$
A direct computation using the operator algebra closed by the
currents of eq. \vcuatro\ shows that $C_A=2$ for ${\rm osp}(2|2)$.
This same calculation yields the values of the metric tensor.
Using these values we can write down the
Sugawara tensor:
$$
\eqalign{
T^J\,=\,{1\over
2k+2}\,&[\,J_+J_-\,+\,J_-J_+\,+\, 2H^2\,-\,2J^2\,
-\,{1\over
2}\,(\,j_{++}\,j_{--}\,-\,\,j_{--}\,j_{++}\,+\,\cr
+&\,j_{+-}\,j_{-+}\,-\,\,j_{-+}\,j_{+-})\,].\cr}
\eqn\vcinco
$$
Taking eq.\vcuatro\ into account, one can obtain the expression of
$T^J\,\,$ in terms of the free fields. After some
calculation one gets:
$$
T^J\,=\,w\partial\chi\,-\,
\bar\psi_+\partial\psi_-\,-\,\bar\psi_-\partial\psi_+
\,-\,{1\over 2}\,(\partial \phi)^2\,-\,
{1\over 2}\,(\partial \varphi)^2.
\eqn\vseis
$$
Notice that the fields $\phi$ and $\varphi$ have vanishing
background charges. It is also evident by inspecting eq.
\vseis\ that the central charge of the ${\rm osp}(2|2)$ WZW model is
zero. This also follows from the fact that
$d_B\,=\,d_F\,=\,4$
for ${\rm osp}(2|2)$(see eq. \zk).
As in the ${\rm osp}(1|2)$ case, we shall denote the ghosts for the
currents $H$ and $J_{\pm}$ by $(\rho_0,\gamma^0)$ and
$(\rho_{\pm},\gamma^{\pm})$, whereas $(\rho_J,\gamma^J)$ and
$(\lambda_{\pm\pm},\eta^{\pm\pm})$ will correspond to the
currents $J$ and $j_{\pm\pm}$ respectively. The topological
${\rm osp}(2|2)$ current system will be realised by combining these
ghosts with two currents whose levels are $k_1=k$ and $k_2=-k-2$. In
order to make contact with the non-critical $N=2$ superstring we
must first improve the energy-momentum tensor. Let us assume that
we deform the total operator $T$ as $T\rightarrow T+\partial{\cal H}$,
where ${\cal H}$ is the total Cartan current along the
$H$-direction. The contributions to ${\cal H}$ of the
currents with levels $k$ and $-k-2$ can be read from the
third equation in \vcuatro. Moreover, using the structure
constants of the ${\rm osp}(2|2)$ algebra, it is easy to compute the ghost
contribution to ${\cal H}$. One gets:
$$
\eqalign{
{\cal H}^{{\rm gh}}\,=&\,
\,\gamma^+\rho_+\,-\,\gamma^-\rho_-\,-\,
{1\over 2}\,\eta^{++}\lambda_{++}\,-\,
{1\over 2}\,\eta^{+-}\lambda_{+-}\,+\,\cr\cr
+&\,{1\over 2}\,\eta^{-+}\lambda_{-+}\,+\,
{1\over 2}\,\eta^{--}\lambda_{--}.\cr}
\eqn\vsiete
$$
The improved energy-momentum tensor of the ghosts can be easily
computed from eq.\vsiete:
$$
\eqalign{
T&_{{\rm improved}}^{{\rm gh}}\,=
\,\rho_0\partial\gamma^0\,+\,
\,\rho_J\partial\gamma^J\,+\,
\gamma^+\partial\rho_+\,+\,2\rho_-\partial\gamma^- \,-\,
\gamma^-\partial\rho_- +\cr\cr
+&\sum_{\alpha = \pm}[
\,{1\over 2}\partial \eta^{+\alpha}\lambda_{+\alpha}
-\,{1\over 2} \eta^{+\alpha}\partial\lambda_{+\alpha}\,+\,
{3\over 2}\partial \eta^{-\alpha}\lambda_{-\alpha}
\,+\,{1\over 2}
\eta^{-\alpha}\partial\lambda_{-\alpha}\,].\cr}
\eqn\vocho
$$
Therefore the conformal weights that the different antighosts
acquire after the deformation are:
$$
\eqalign{
&\Delta (\rho_0)\,=\,\Delta (\rho_J)\,=\,1
\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\rho_+)\,=\,0
\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\rho_-)\,=\,2\cr\cr
&\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\lambda_{+\pm})\,=\,{1\over 2}
\,\,\,\,\,\,\,\,\,\,\,\,
\Delta (\lambda_{-\pm})\,=\,{3\over 2}.\cr}
\eqn\vnueve
$$
Moreover, from eqs. \vseis\ and \vcuatro, one can obtain the improved
energy-momentum tensor in the current sector. If we label with the
indices $1$ and $2$ the free fields coming from the two currents, it is
straightforward to arrive at the result:
$$
\eqalign{
T&_{{\rm improved}}^{\rm J}\,=\,
\sum_{i=1,2}[\,
-{1\over 2}\,(\partial\varphi_i)^2\,
-\,{1\over 2}\,(\partial\phi_i)^2\,]\,+\,
{i\over 2}\,\sqrt{2k+2}\,\partial^2\phi_1\,-\,
{1\over 2}\,\sqrt{2k+2}\,\partial^2\phi_2\,-\cr\cr
-&\sum_{i=1,2}\,[\partial w^i\chi^i\,
+\, {1\over 2}\bar \psi^i_+\partial\psi^i_-\,
+\, {1\over 2}\psi^i_-\partial\bar\psi^i_+\,
+\, {1\over 2}\bar \psi^i_-\partial\psi^i_+\,
+\, {1\over 2}\psi^i_+\partial\bar\psi^i_-\,].\cr}
\eqn\treinta
$$
Let us now see how one can pair ghost fields from eq.\vocho\ with
fields in eq. \treinta\ in such a way that, after the reduction, we are
left with the field content of the non-critical $N=2$ superstring.
Indeed, as in the ${\rm osp}(1|2)$ case, we can pair the systems
$(\rho_0,\gamma^0)$ and $(\rho_+,\gamma^+)$ with
$(w^1,\chi^1)$ and $(w^2,\chi^2)$. Moreover we can form a quartet with
the ghosts $(\lambda_{\pm\pm}, \eta^{\pm\pm})$ and the
$({1\over 2},{1\over 2})$ anticommuting system formed from the Majorana
fermions ${1\over 2}(\psi^1_{\pm}+\bar \psi^1_{\mp})$ and
${1\over 2}(\psi^2_{\pm}+\bar \psi^2_{\mp})$. The remaining fields can
be assigned to matter and Liouville degrees of freedom. Let us first
consider the bosonic fields. If the labels M and L refer matter and
Liouville fields, we can define:
$$
\eqalign{
\phi_M\,=&\,{1\over\sqrt 2}\,(\phi_1\,+\,i\varphi_1)
\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
\bar\phi_M\,=\,{1\over\sqrt
2}\,(\phi_1\,-\,i\varphi_1)\cr\cr
\phi_L\,=&\,{1\over\sqrt 2}
\,(\phi_2\,+\,i\varphi_2)
\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
\bar\phi_L\,=\,{1\over\sqrt 2}\,
(\phi_2\,-\,i\varphi_2).\cr}
\eqn\tuno
$$
Similarly we can define the Dirac fermionic fields:
$$
\eqalign{
\psi_M\,=&\,{i\over 2}\,[\psi^1_+\,-\,\bar\psi^1_-\,+\,
i(\psi^1_-\,-\,\bar\psi^1_+)\,]
\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
\bar\psi_M\,=\,{i\over 2}\,[\psi^1_+\,-\,\bar\psi^1_-\,-\,
i(\psi^1_-\,-\,\bar\psi^1_+)\,]\cr\cr
\psi_L\,=&\,{i\over 2}\,[\psi^2_+\,-\,\bar\psi^2_-\,+\,
i(\psi^2_-\,-\,\bar\psi^2_+)\,]
\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
\bar\psi_L\,=\,{i\over 2}\,[\psi^2_+\,-\,\bar\psi^2_-\,-\,
i(\psi^2_-\,-\,\bar\psi^2_+)\,].\cr\cr}
\eqn\tdos
$$
It should be noticed that $\psi_M$ and $\bar\psi_M$
($\psi_L$ and $\bar\psi_L$) are the components of the fermionic fields
used to represent the ${\rm osp}(2|2)$ current at level $k_1$ (respectively
$k_2$) that are not affected by the reduction described above.
Splitting the reduced energy-momentum tensor as:
$$
T_{{\rm reduced}}\,=\,T_{{\rm reduced}}^{\rm M+L}\,+\,
T_{{\rm reduced}}^{\rm gh},
\eqn\ttres
$$
we can easily write down the matter + Liouville contributions.
Using the definitions \tuno\ and \tdos, one arrives at:
$$
\eqalign{
T_{{\rm reduced}}^{\rm M+L}\,=&\,
-\partial\phi_M\partial\bar\phi_M\,
-\partial\phi_L\partial\bar\phi_L\,+\,
{\sqrt{k+1}\over 2}\,
(i\partial^2\,(\phi_M+\bar\phi_M)\,-\,
\partial^2\,(\phi_L+\bar\phi_L))-\cr\cr
-&{1\over 2}\,\bar\psi_M\partial\psi_M
-{1\over 2}\,\psi_M\partial\bar\psi_M-
{1\over 2}\,\bar\psi_L\partial\psi_L
-{1\over 2}\,\psi_L\partial\bar\psi_L.\cr\cr}
\eqn\tcuatro
$$
The ghost part of $T_{{\rm reduced}}$ can be read form eq. \vocho.
Relabelling the ghost fields unaffected by the reduction as
$\rho_-\rightarrow b$, $\gamma^-\rightarrow c$,
$\rho_J\rightarrow \tilde b$, $\gamma^J\rightarrow \tilde c$,
$\lambda_{-\pm}\rightarrow \beta_{\pm}$ and
$\eta^{-\pm}\rightarrow \gamma^{\pm}$, one can finally write
$T_{{\rm reduced}}^{\rm gh}$ in the form:
$$
T_{{\rm reduced}}^{\rm gh}\,=\,
2b\partial c\,-\,c\partial b\,+
\,{3\over 2}\,\partial\gamma^+
\beta_+\,+\,{1\over 2}\,\gamma^+\partial\beta_+\,+\,
\,{3\over 2}\,\partial\gamma^-
\beta_-\,+\,{1\over 2}\,\gamma^-\partial\beta_-\,+\,
\tilde b\partial\tilde c,
\eqn\tcinco
$$
which indeed coincides with the energy-momentum tensor for the
ghosts of the $N=2$ string. Notice that the contribution of this $N=2$
ghost system to the Virasoro central charge is $-6$. Moreover the
matter central charge $c_M$ and the ${\rm osp}(2|2)$ level $k$ are now
related as:
$$
c_M\,=\,-6k-3,
\eqn\tseis
$$
in complete agreement with ref. [\bershadsky]
\chapter{Summary and conclusions.}
In this paper we have analysed the topological
structure associated with the Ka\v c--Moody symmetry
based on an arbitrary Lie superalgebra ${\cal G}$. The resulting
topological algebra turns out to
be an extension of the twisted $N=2$ superconformal
algebra [\LVW,\EY] by a BRST doublet of spin 3
operators, upon whose introduction the algebra
closes linearly.
Our method consists in introducing a ghost sector that
allows to define a BRST cohomology, in such a
way that all operators appear as BRST doublets. The
resulting theory has been interpreted as a
topological sigma model for supergroup manifolds,
\ie, as a ${\cal G}/{\cal G}$ coset.
By means of free field representations we have related the ${\rm osp}(1|2)$
and
${\rm osp}(2|2)$ models with the $N=1$ and $N=2$ non-critical superstrings,
respectively. We have established that, after performing a suitable
deformation of the coset model, one can identify the matter, Liouville
and ghost sectors of the coset with those of the
corresponding string theory. Moreover the
remaining degrees of freedom can be accommodated in topological doublets.
In this paper we have restricted ourselves to comparing the field content
of the deformed coset models and non-critical superstrings. It remains
to see whether or not the BRST cohomologies of both models are related.
For bosonic
$W_N$-strings this relation has been shown to exist
[\yank,\hu,\sadov]
and thus one expects a similar result for the supersymmetric string
models (see ref. [\yu] for an analysis of the ${\rm osp}(1|2)$ case).
There are many other
questions that remain open in the relation between the
${\cal G}/ {\cal G}$ cosets and string theories. For example, one
would expect to have a current algebra prescription to compute
correlation functions in non-critical string theories. Another
problem that, in our opinion, deserves future investigation, is the
relation between the ${\cal G}/{\cal G}$ BRST symmetry and the
topological symmetry of non-critical strings discovered in ref.
\REF\gato{B. Gato-Rivera and A.M.
Semikhatov\journalPhys. Lett.&B293(92)72 \journalNucl. Phys.&B408(93)133.}
\REF\BLNW{M. Bershadsky, W. Lerche, D. Nemechansky and N.
Warner\journalNucl. Phys.&B401(93)304.}
[\gato] (see also [\BLNW]).
The ${\cal G}/{\cal G}$ coset models can also be regarded as
models of topological matter. The coupling of this
${\cal G}/ {\cal G}$ matter to topological gravity gives rise to
a topological string model. A prescription for this coupling was given
in ref. [\getzler] at the level of the operator algebra. The extended
character of the BRST algebra plays an important role in this analysis.
It would be interesting to understand this coupling within a lagrangian
formalism. This could possibly shed light on the nature and
implications of the BRST symmetry of current algebras.
\ack
The authors would like to thank
J. M. F. Labastida and J. S\'anchez Guill\'en for encouragement.
Discussions with J. S\'anchez de Santos, P. M. Llatas
and L. Ferreira are gratefully acknowledged. J. M. I. is
thankful to the Theory Group of the Physics Department
of Brandeis University for hospitality,
and to Ministerio de
Educaci\'on y Ciencia (Spain) for financial support.
This work was supported in part by DGICYT under
grant PB90-0772, and by CICYT under grants AEN90-0035 and AEN93-0729.
\refout
\end
|
\section{Introduction}
\label{intro}
The production of photons in high momentum transfer collisions is useful for a
variety of reasons \cite{JFO}. The single photon invariant cross section is
one of
several important sources of information on the gluon distribution through the
role played by the subprocess $qg\rightarrow \gamma q$ \cite{OT}. The photon
plus jet
cross section may also provide additional constraints as the quantity of such
data is increased. Photon pair production has already played a role in
confirming the quark charge assignments and is one of many large momentum
transfer processes used to check the QCD-based description of high-p$_T$
scattering \cite{BBF}. Most recently, photon pair production has received
attention
because of its role in the search for the Higgs boson in the intermediate mass
region \cite{GKW}. The signature provided by the decay $H\rightarrow \gamma
\gamma$ is
particularly clean, enabling one, in principle, to detect the relatively rare
processes of Higgs production. However, a crucial element in this search is to
understand the two-photon background that results from conventional scattering
processes \cite{BO}. In addition, it has been suggested that the observation of
the
production of photon pairs may serve as a useful signal for new physics
processes\cite{Rizzo}.
The purpose of this paper is to present predictions for the event structure in
two photon events. In particular, we will focus on events with one additional
jet in the final state. Such events can be conveniently analyzed in terms of
several angular distributions and the structure of a two-dimensional Dalitz
plot. In addition, the presence of the jet allows one to define three classes
of events depending on whether the jet has the most, second most, or least
transverse energy when compared to the two photons. A similar analysis has been
proposed for events with one photon and two jets in the final state\cite{KO}.
The plan of the paper is as follows. In Sec.\ \ref{sec:program} a brief
description of the program will be given. In Sec.\ \ref{sec:plots} the
characteristics of the two photon plus jet final states will be presented for a
set of cuts which could be appropriate for either the CDF or D0 experiments.
Our summary and conclusions will be presented in Sec.\ \ref{sec:summary}.
\section{${\cal O}(\alpha^2 \alpha_s)$ Calculation}
\label{sec:program}
The program used for this analysis has been described in Ref.\ \cite{BOO},
so only a brief summary will be presented here. To order ${\cal O}(\alpha^2)$
the production of photon pairs proceeds via $q\bar q \rightarrow \gamma
\gamma$. In addition, there are contributions from final states with a
single photon and a jet where the jet gives rise to a photon. This is
estimated by including the $q\bar q \rightarrow \gamma g \text{ and }
qg\rightarrow \gamma q$ subprocesses convoluted with the appropriate
photon fragmentation function. The latter is formally of ${\cal O}(\alpha
/\alpha_s)$ which yields an overall factor of $\alpha^2$ when combined
with the factor of $\alpha \alpha_s$ coming from the subprocess. Similarly,
one can have final states with two jets, each of which fragments to a photon
plus hadrons. This contribution is given by the sum of all $2\rightarrow2$
parton-parton scattering subprocesses convoluted with two fragmentation
functions. These last two contributions are referred to as single and
double bremsstrahlung processes, respectively. Although formally of ${\cal O}
(\alpha^2 \alpha_s^2)$, one can also have significant contributions from the
subprocess $gg\rightarrow \gamma \gamma$ at high energies where small values
of $x$ become important since the gluon distribution is large enough there to
compensate for the suppression from the running coupling.
The ${\cal O}(\alpha^2 \alpha_s)$ next-to-leading-logarithm calculation
includes contributions from the one-loop corrections to $q\bar q \rightarrow
\gamma \gamma$ and the $2\rightarrow 3$ subprocesses $q\bar q \rightarrow
\gamma \gamma g \text{ and } qg\rightarrow \gamma \gamma q$. The actual
calculation is performed using a combination of Monte Carlo and analytic
integration techniques. The regions of phase space corresponding to soft or
collinear singularities are isolated by the use of cutoffs, $\delta_s
\text{ and } \delta_c$, applied to the Mandelstam variables for the
$2\rightarrow 3$ subprocesses. The squared matrix elements are integrated over
the singular regions using dimensional regularization. The soft singularities
cancel with corresponding terms from the one-loop corrections. Finally, the
collinear singularities are factorized using the ${\rm \overline{MS}}$ scheme
and
absorbed into the corresponding distribution or fragmentation function.
All remaining integrations are performed via Monte Carlo. When properly
executed, the cutoff dependence in such a calculation cancels between the
two-body and three-body contributions which is appropriate since the cutoffs
merely serve to mark where one switches from a Monte Carlo to an
analytic integration technique. Additional details can be found in Ref.\
\cite{BOO}.
The calculation outlined here includes contributions from all ${\cal
O}(\alpha^2 \alpha_s)$ subprocesses where the photons both are part of the
hard scatter. There are also fragmentation contributions where the hard
scattering process is calculated to ${\cal O}(\alpha_s^3)$ and two of the
jets fragment to photons. These contributions have not been incorporated into
the calculation. However, for collider energies isolation cuts are employed to
help reduce the hadronic background to the photon signals. These isolation cuts
greatly reduce the fragmentation contributions. Another point to remember is
that this calculation is
next-to-leading-order for observables involving the photon pairs. However, for
the two photon plus jet final states only the $2\rightarrow 3$ subprocesses
contribute, so this represents a leading order calculation for that case.
There are, for example, corrections from $2\rightarrow 4$ subprocesses where
one of the final jets is not detected. These contributions are beyond the scope
of this calculation.
\section{Predictions}
\label{sec:plots}
Predictions for two photon and two photon plus jet final states were
generated using the program described in the previous section. The CTEQ2M
distributions \cite{CTQLNG} were used with the factorization and
renormalization scales chosen to be the maximum $E_T$ in each event. A
minimum transverse momentum cut of 10 GeV is applied to the photons and the
jet. In addition, the pseudorapidities of the two photons are required to
satisfy $\vert \eta \vert
\leq 1$ and that of the jet to satisfy $\vert \eta \vert \leq 3$.
Both photons are required to be isolated, {\it i.e.}, to have less than
4 GeV of additional energy in a cone of radius $R = \sqrt{\Delta \eta^2
+\Delta \phi^2}$ of 0.7 centered on the photon direction. Effectively, this
just requires the two photons and the jet to be separated by more than
$\Delta R=0.7$. Such a cut greatly reduces contributions from subprocesses
involving photon fragmentation functions. In the experimental analysis the
isolation cut also discriminates against jets fragmenting into a leading
$\pi^0 \ {\rm or\ }\eta$, and thus greatly decreases the background to the
true two photon signal. These cuts are typical of what would be appropriate
for an analysis by either the D0 or CDF groups.
In Fig.\ \ref{etxsec} the cross section is shown for two photon and two photon
plus jet production as a function of the total transverse energy
($E_T^{\gamma \gamma}$) of the two photons. The photon and jet (if present)
are subject to the cuts
described above. The top curve shows the total cross section while the bottom
curve shows the cross section with a jet satisfying the above cuts.
The fraction of the cross section coming from events
with a jet in the final state that satisfies these cuts varies from about 60\%
for $E_T^{\gamma \gamma} = 100$ GeV to about 30\% at the upper end of the
curve. This ratio is larger than might have been expected since the rapidity
interval for the accompanying jet is much larger than that for the photons.
In Fig. \
\ref{subproc} the fractions of the two photon plus jet cross section coming
from the $q\bar q\ {\rm and\ } qg$ subprocesses are shown versus the total
transverse energy in the final state, denoted by $\Sigma E_T$. At low $\Sigma
E_T$ the $qg$ subprocess dominates while the $q\bar q$ subprocess dominates
at high $\Sigma E_T$ values. This is primarily due to the relative sizes of
the gluon and quark distributions at low and high values of $x$.
For those events with a jet in the final state, the fractions of the cross
section are plotted
in Fig. \ \ref{etjet} according to whether the jet has the largest, second
largest, or least $E_T$ among the two photons and the jet. It is useful to
classify the events in this manner since the jet is the one object
distinguishable form the others. The photon was used in this manner in a
similar analysis of photon plus two jet events in Ref.\ \cite{KO}. The results
in this figure show that at high $\Sigma E_T$ it is most likely that the jet
will have less $E_T$ than either of the photons, while at lower values it is
more likely that the jet will have more $E_T$ than one or both of the photons.
For events with two photons and a jet in the final state, four kinematic
variables in addition to the three-body center-of-mass energy must be specified
in order to determine the final state configuration. It is convenient to use
scaled energy variables for two of these. In the parton-parton center-of-mass
frame the energies of the final state particles can be labelled as $E_3,
E_4, \ {\rm and \ } E_5$ in decreasing order of energy. The energy fractions
$x_i
=2 E_i/E_{tot}$ can then be constructed, with $E_{tot}=E_3+E_4+E_5$. Hence,
$x_3 + x_4 + x_5 =2$. The variable $x_3$ satisfies $\frac 2 3 \leq x_3 \leq 1$,
where the lower limit corresponds to the symmetric ``Mercedes Benz''
configuration with all three $x_i's$ being equal. As $x_3$ tends towards one,
the vectors of particle 4 and 5 become collinear and
opposite to that of particle 3. The allowed range for $x_4$ is $0.5 \leq
x_4 \leq 1$ where the lower limit corresponds to $x_4=x_5$ with particle 4 and
5 being collinear while the upper limit corresponds to $x_5=0 \text{ and }
x_3+x_4=1$. Only two of the $x_i's$ are independent; these will be taken to be
$x_3 \text{ and } x_4$. Two angles are needed to complete the description of
the final state. In the parton-parton
center-of-mass frame, let $\theta^*$ be the angle between the direction of the
incoming proton and the
direction of the jet or photon with energy $E_3$. The plane containing
the beam and this jet or photon will be referred to as the scattering plane.
The second angle is $\psi$ which is the angle between the normal to the
scattering plane and the normal to the plane containing the remaining jet or
photons.
The main point of this analysis is that examination of various distributions
in terms of these kinematic variables may provide useful information concerning
the accuracy of the current QCD-based description of two photon production.
For example, some lego plots of $x_3 \text{ versus } x_4$ for events where the
jet is the most energetic are shown in Fig.\ \ref{jet1-60}, \ref{jet1-200}, and
\ref{jet1-400}. In these cases the
two photons recoil against the jet. The $p_T, \eta,$\ and isolation cuts
described above have been applied. In Fig.\ \ref{jet1-60} a cut requiring the
total $E_T$ to be greater than 60 GeV has been applied. No obvious structure
is visible. However, any structure is hidden because the total $E_T$ cut is
low with respect to the individual minimum $E_T$ requirements. In Fig.\
\ref{jet1-200} the cut on the total $E_T$ has been raised to 200 GeV and in
Fig.\
\ref{jet1-400} to 400 GeV. As the minimum for the total $E_T$ is raised an
enhancement in the region $x_3, x_4 \rightarrow 1$ begins to appear, reflecting
the existence of a pole in the matrix element at $x_3=x_4=1$. In this region
one of the photons becomes soft, {\it i.e.,} approaches the minimum $E_T$
value of 10 GeV. For high values of the total $E_T$ it is expected that the
majority of events in this class would have a high $E_T$ jet, a slightly less
energetic photon, and a relatively soft second photon.
The lego plots for the case where the jet is the second most energetic show a
similar structure. A related, but somewhat different, pattern emerges when the
jet has less energy than either of the photons. For this configuration
the jet and the photon with the least energy of the pair recoil against the
most energetic photon. Again, a strong pole structure becomes increasingly
evident for progressively higher values of total $E_T$. However, as shown in
Fig.\ \ref{jet3-400}, the shape
of the emerging pole structure is somewhat broader in this case as compared to
the case where the jet is the most energetic. This can be ascribed to the fact
that the subprocess $qg \rightarrow \gamma \gamma q$ does not have a
soft singularity when the quark energy goes to zero, but it does when one of
the photon energies goes to zero. This can be further illustrated by comparing
Fig.\ \ref{qqbar} and Fig.\ \ref{qg} which show the contributions to Fig.\
\ref{jet3-400} from the $q\bar q \rightarrow \gamma \gamma g \text{ and }
qg\rightarrow \gamma \gamma q$ subprocesses, respectively.
We have examined the $\cos(\theta^*) \text{ and } \psi$
distributions for the three cases where the jet has the most, second most, or
least energy with minimum total $E_T$ cut of 200 GeV applied. The $\psi$
distributions are essentially the same for all three cases while some minor
differences are apparent in the $\cos(\theta^*)$ distributions. However, the
exact shapes are rather cut dependent and so have not been shown. The
appropriate distributions can easily be generated once the appropriate cuts
have been specified.
Several two photon events have been observed by the CDF Collaboration with
values of total $E_T$ exceeding 400 GeV \cite{CDF}. Two of the three events
have a jet associated with the two photons. In one event the jet
is the most energetic and for the other it is the least energetic. Based on
the previous discussion one might expect that the photon or jet with the least
energy might have an $E_T$ value near the minimum allowed, {\it i.e.,} that the
event would be in the region of the corresponding lego plot where the
enhancement due to the pole occurs. Instead, the energies are closer than would
have been anticipated. The unusual topologies of these events may simply be
due to statistical fluctuations, but the continued presence of a large number
of such events as the data sample grows could signal the need for additional
sources of high mass photon pairs in the calculation.
\section{Summary and Conclusions}
\label{sec:summary}
As preparations for the next generation of colliders continues, interest in
the $\gamma \gamma$ channel has remained high since it can be used to search
for the intermediate mass Higgs boson. It is, therefore, necessary to
understand the conventional QCD production mechanisms for photon pairs in
order to be able to provide reliable background estimates. The study of the
event structure of two photon plus jet events using the distributions suggested
in this paper may well provide useful tests of our understanding.
|
\subsection*{1. Form factors of heavy-to-light meson transitions}
The exclusive semileptonic $B$ decays to $\pi$ and $\rho$ play a prime role in
the measurement of $V_{ub}$. Let us consider, for example, the spectrum of
${\bar {B^0}} \to \pi^+ \mu^- {\bar\nu}$:
\begin{equation}
{d \Gamma ({\bar {B^0}} \to \pi^+ \mu^- {\bar\nu})\over d q^2}=
{G_F^2 \over 24 \pi^3} |V_{ub}|^2 |F_1^{B\pi}(q^2)|^2
|\vec p'_\pi(q^2)|^3 \label{dg}
\end{equation}
where $\vec p'_\pi(q^2)$ is the pion three-momentum (at fixed $q^2$) in the
$B$ meson rest frame. It is clear that a measurement of
$d \Gamma\over d q^2$ would provide us with $V_{ub}$ once the form factor
$F_1^{B\pi}(q^2)$, defined by the hadronic matrix element ($P=\pi$):
\begin{equation}
< P (p^\prime) | \bar q \gamma_\mu b | B(p)> =
(p + p^\prime)_\mu F_1 (q^2) +
{ M^2_B - M^2_P \over q^2 }
[F_0(q^2) - F_1(q^2)] q_\mu\hskip 2pt \label{sp0}
\end{equation}
\noindent
($q=p-p'$, $F_0(0)=F_1(0)$) is known in the whole accessible range of $q^2$:
$0\le q^2\le q^2_{max}=(M_B-M_\pi)^2$.
An equation analogous to (\ref{dg}) holds for
${\bar {B^0}} \to \rho^+ \mu^- {\bar\nu}$ in terms of the form factors
$V$ and $A_i$ defined by the matrix element ($P^*=\rho$):
\begin{eqnarray}
<P^* (p^\prime, \eta) | \bar q \gamma_\mu (1- \gamma_5) b
| B(p)> &=&
\epsilon_{\mu \nu \rho \sigma}
\eta^{* \nu} p^\rho p^{\prime \sigma} { 2 V(q^2) \over M_B + M_{P^*}}
- i (M_{B} + M_{P^*}) A_1 (q^2) \eta_{\mu}^{\ast} \;
\nonumber \\
+ i {A_2 (q^2)
\over M_{B} + M_{P^*}} (\eta^{\ast}\cdot p) (p + p^\prime)_\mu
&+& i \;(\eta^{\ast}\cdot p) {2 M_{P^*} \over q^2} q_\mu
\big(A_3 (q^2) - A_0 (q^2)\big)
\label{sp1}
\end{eqnarray}
\noindent ($A_3(0)=A_0(0)$ and
$A_3 (q^2)\;=\;{M_{B}+M_{P^*} \over 2 M_{P^*}} A_1 (q^2) -
{M_{B}-M_{P^*} \over 2 M_{P^*}} A_2 (q^2)\; \;$).
Model independent relations can be derived for $F_i$, $V$ and $A_i$ in the
infinite heavy quark mass limit at the point of
zero recoil ($q_{max}^2$) where $\pi$ and $\rho$ are at rest in the $B-$meson
rest frame \cite{IW1}.
For example, when $m_b \to \infty$ the following relations can be
worked out for $F_1$ and $F_0$:
\begin{eqnarray}
F_1^{B\pi}(q_{max}^2)& \simeq &{g \over 2 f_\pi} {\hat F}
{\sqrt {M_B} \over \Delta +
M_\pi} \label{f1}\\
F_0^{B\pi}(q_{max}^2)& \simeq & - {\hat F \over f_\pi } {1 \over \sqrt {M_B}}
+ {\cal O} (M_\pi) \; ; \label{f0}
\end{eqnarray}
eq.(\ref{f1}) describes the dominance of the $B^*$ pole for $F_1$ at zero
recoil ( in the limit $m_b \to \infty$
$\hat F$ is related to $f_B$ and $f_{B^*}$, $g$ is the rescaled
$B^*B\pi$ strong coupling, and $\Delta=M_{B^*} - M_B$); eq.(\ref{f0}) is
the Callan-Treiman relation valid in the chiral limit.
The above scaling relations, that could be used, e.g., to relate
$B \to \pi \ell \nu$ to $D \to \pi \ell \nu$ at zero recoil, are not sufficient
to describe the form factors in the physical range of
transferred momentum; therefore, a dynamical calculation
based on QCD is required for $F_i(q^2)$, $V(q^2)$ and $A_i(q^2)$.
The method of QCD sum rules \cite{SVZ}
is a fully relativistic field-theoretical approach
incorporating fundamental features of QCD, such as perturbative asymptotic
freedom and nonperturbative quark and gluon condensation.
This method allows us, by analyzing three-point correlators of
quark currents, to compute the form factors from zero
to quite large values of $q^2$; in
this respect, the method complements lattice QCD, where B meson form
factors, extrapolated from the charm quark mass, are
computed in the region near $q^2_{max}$
\cite{APE,UKQCD,Gu}.
Several QCD sum rules calculations of $F_1^{B \pi}$ can be found in the
literature \cite{Ball}; a calculation of both $F_1$ and $F_0$ has been
performed in ref.\cite{Col1} in the limit $m_b \to \infty$. The result
for $F_1^{B \pi}(q^2)$, depicted in
fig.1 and common to other QCD sum rules analyses,
supports the simple pole model:
$F_1^{B \pi}(q^2)= {[0.24 \pm 0.02] /( 1- {q^2 \over M^2_{B^*}})}$;
on the other hand $F_0^{B \pi}(q^2)$ increases slowly with $q^2$.
The feature of $F_0$ of
being nearly independent of $q^2$ has been confirmed by a
calculation, at finite $m_b$, in the channel $B \to K$ (fig.2a)
\cite{Col2}.
\vskip 6.truecm\noindent
\special{psfile="f0_bp.ps" hoffset=90 voffset=-50 hscale=65 vscale=65}
\par
\centerline{Fig.1: Form factors $F_1^{B \pi}$
(continuous line) and $F_0^{B \pi}$ (dashed line).}
The computed $q^2$ dependence of $F_0^{B \pi}(q^2)$
must be compared with the
expectation based on the hypothesis of
the dominance of the nearest singularity in the $t-$
channel, assumed in a number of models \cite{mod}:
the nearest pole contributing to
$F_0^{B \pi (B K)}$ is the $0^+$ $b \bar u$ ($b \bar s$) state
with mass in a range near $6 \; GeV$ (in the BSW model the
value $M_{(b {\bar s})}(0^+)=5.89 \; GeV$ is used);
on the other hand, a fit of the obtained
$F_0^{B \pi}(q^2)$ and
$F_0^{B K}(q^2)$ to a simple pole
formula can be performed provided that $M_P \ge 7 - 7.5 \; GeV$.
This different $q^2$ behavior of
$F_1^{B \pi}(q^2)$ and $F_0^{B \pi}(q^2)$
has been observed also in lattice QCD \cite{UKQCD, Gu}. Moreover, a
different functional dependence is expected if one considers that the
scaling laws with the heavy mass in eqs.(\ref{f1},\ref{f0}) are compatible
with the relation $F_1(0)=F_0(0)$ if the $q^2$ dependence is, e.g., of the
type:
$F_i(q^2)= {F_i(0)/ (1- {q^2 \over M^2_i})^{n_i}}$, with
$n_1=n_0+1$.
Deviations from the single pole model have been observed also for the
axial form factors
$A_1^{B \rho}$ and $A_2^{B \rho}$, that turn out to be rather flat in $q^2$
(see the first article in ref.\cite{Ball}); on the other hand,
$V^{B \rho}$ can be fitted with a polar formula, the pole given by $B^*$
\footnote {
A steeper increase of $V$ compared to $A_1$ is obtained
also in ref.\cite{Ali},
but the slopes are different with respect to Ball's results in \cite{Ball}.}.
As for the last form factor in eq.(\ref{sp1}), $A_0$,
the calculation both in the channels $B \to \rho$ and $B \to K^*$
\cite{Col2,Col3} shows that it also increases like a pole, with the
pole mass compatible with the mass of $B$ (or $B_s$) as expected by the
nearest-resonance dominance hypothesis (fig.2b). Interesting enough,
the relation $A_0(0) \simeq F_0(0)$ is obtained.
To summarize the results from QCD Sum rules analyses, the following scenario
emerges for the transitions $B \to \pi, \rho$ ( $B \to K, K^*$):
$F_1$, $V$ and $A_0$ following a polar dependence in $q^2$,
$F_0$, $A_1$ and $A_2$ rather flat in $q^2$,
$A_0(0)=F_0(0)=F_1(0)$.
It is worth reminding that such results are obtained after an involved
analytic and numerical analysis, independent for each one of the above
form factors.
\vskip 6.5 truecm\par
\special{psfile="f0_bk.ps" hoffset=-60 voffset=-180 hscale=65 vscale=65}
\special{psfile="a0_bks.ps" hoffset=+160 voffset=-180 hscale=65 vscale=65}
\par
\vskip 0.5 cm
\centerline{ Fig.2: Form factors $F_0^{B K}$ (a) and
$A_0^{B K^*}$ (b).}
One could wonder whether
QCD sum rules results suggest the existence of relations
among the form factors governing the transitions of heavy mesons to light
mesons. For semileptonic decays where both the initial
and the final meson contains one heavy quark, such relations can be derived
in the limit $m_Q \to \infty$:
they connect the six form factors as in
(\ref{sp0},\ref{sp1}) to the Isgur-Wise function \cite{IW} incorporating
the nonperturbative dynamics of the light degrees of freedom.
It is intriguing that relations among heavy-to-light form factors have been
obtained in a constituent quark model by B.Stech \cite{Stech},
assuming that the spectator particle retains its
momentum and spin before the hadronization, and that
in the rest frame of the hadron the constituent quarks
have the off-shell energy close to the constituent mass.
Under these hypotheses the following equations can be written for
$B \to \pi, \rho$ form factors
\footnote{ A dependence of $F_1$ on the mass of the final particle has
to be taken into account since there is no spin symmetry in the
final state.}:
\begin{eqnarray}
&&F_0(q^2)=\left(1-\frac{q^2}{m_B^2-m^2_\pi}\right)
F_1(q^2) \label{r1} \\
&&V(q^2)=\left(1+\frac{m_\rho}{m_B}\right)
F_1(q^2) \label{r2} \\
&&A_1(q^2)=\frac{1+\frac{m^2_\rho}{m^2_B}}
{1+\frac{m_\rho}{m_B}}\left
(1-\frac{q^2}{m^2_B+m^2_\rho}\right)F_1(q^2) \label{r3} \\
&&A_2(q^2)=\left(1+\frac{m_\rho}{m_B}\right)\left(
1-\frac{2m_\rho/(m_B+m_\rho)}{1-q^2/(m_B+m_\rho)^2}\right)
F_1(q^2) \label{r4} \\
&&A_0(q^2)=F_1(q^2). \label{r5}
\end{eqnarray}
The above relations are very similar to the relations holding
for heavy-to-heavy transitions, e.g. $B \to D, D^*$.
QCD sum rules results seem to confirm them.
As a matter of fact, for a polar dependence of
$F_1(q^2)$ the above relations suggest that both
$F_0$ and $A_1$ should be nearly constant in $q^2$, and that
$A_0$ should be equal to $F_1$.
Of course, more work is needed to put
equations (\ref{r1}-\ref{r5}) on the same theoretical grounds of
the relations among the form factors of heavy-to-heavy transitions.
\subsection*{2. Tests of factorization for color suppressed B decays}
Semileptonic form factors are useful not only to predict
semileptonic BR's, but
also to compute nonleptonic two-body decay rates if the factorization
approximation is adopted. In particular, for color suppressed transitions
$B \to K^{(*)} \; J/\Psi$ and
$B \to K^{(*)} \; \eta_c$, only the heavy-to-light form factors are needed.
The decays
$B \to K^{(*)} \; J/\Psi $ have been analyzed in
\cite{Gourdin} to constrain the semileptonic
$B \to K^*, K $ form factors using data on the
longitudinal polarization of the final
particles in the decay $B \to K^* \; J/\Psi$:
$\rho_L=\Gamma(B \to K^* \; J/\Psi)_{LL}/\Gamma(B \to K^* \; J/\Psi)
=0.84 \pm 0.06\pm 0.08$, and on the ratio
$R_{J/\Psi}=\Gamma(B \to K^* \; J/\Psi)/\Gamma(B \to K \; J/\Psi)
=1.64 \pm 0.34$ \cite{Hon}.
In the same spirit, the decays $B \to K^{(*)} \; \eta_c $ are interesting
since they could help
in testing the factorization scheme and the accuracy of the computed
hadronic quantities \cite{Col2}.
To predict the decay rates $B \to K^{(*)} \; \eta_c $, besides
$F_0^{B K}(q^2)$ and $A_0^{B K^*}(q^2)$ we need the leptonic
constant $f_{\eta_c}$. Together with $f_{\eta'_{c}}$, $f_{\eta_c}$
can be obtained by QCD sum rules considering the two-point function
\begin{eqnarray}
\psi_{5}(q) = i \int d^{4}x \; e^{iqx}\; < 0| T (\partial^{\mu} A_{\mu} (x)
\partial^{\nu} A_{\nu}^{\dagger} (0)) |0 > \;
\end{eqnarray}
($\partial ^{\mu} A_{\mu} (x) = 2 m_{c} : \bar{c}(x) i \gamma_{5}c(x):$)
that is known in perturbative QCD to two-loop
order, including also the leading $D=4$ non-perturbative
term in the Operator Product Expansion.
Exploiting two different types
of QCD sum rules, viz. Hilbert transforms at $Q^{2} = 0$, and Laplace
transforms, we get \cite{Col2}:
\begin{equation}
f_{\eta_{c}} \simeq 301 - 326 \; \mbox{MeV},\;\;\;\;\;\;\;\;
f_{\eta'_{c}} \simeq 231 - 255 \; \mbox{MeV},\;
\end{equation}
\begin{equation}
f_{\eta_{c}} \simeq 292 - 310 \; \mbox{MeV},\;\;\;\;\;\;\;\;
f_{\eta'_{c}} \simeq 247 - 269 \; \mbox{MeV}\;,
\end{equation}
respectively. These results have been obtained by varying the parameters
in the ranges
dictated by the gluon condensate and quark-mass analyses, and using
$m_{c} = 1.46 \pm 0.07 \; \mbox{GeV}$,
$\Lambda_{QCD} = 200 - 300 \; \mbox{MeV}$, with the constraint that
$M_{\eta_{c}}$ and $M_{\eta'_{c}}=3595\pm5\;MeV$ are
correctly reproduced by the sum rules.
Combining the predictions from the Hilbert and Laplace method we obtain:
\begin{equation}
f_{\eta_{c}} = 309 \pm 17 \; \mbox{MeV},\;\;\;\;\;\;\;\;
f_{\eta'_{c}} = 250 \pm 19 \; \mbox{MeV},\;\;\;\;\;\;\;\;
\frac{f_{\eta'_{c}}} {f_{\eta_{c}}} = 0.8 \pm 0.1 \; ,
\label{pred}
\end{equation}
\begin{equation}
{f_{\eta_c} \over f_{J/\psi}} = 0.81 \pm 0.05,\;\;\;\;\;\;\;\;
{f_{\eta^\prime_c} \over f_{\Psi^\prime}} = 0.88 \pm 0.08 .
\label{pred1}
\end{equation}
In (\ref{pred1}) the experimental values: $f_{J/\Psi}=384 \pm 14 \; MeV$ and
$f_{\Psi^\prime}=282 \pm 14 \; MeV$ have been used.
In the constituent quark model the leptonic constants of the charmonium system
can be expressed in terms of the $c \bar c$ wave function at the origin
$\Psi(0)$:
\begin{equation}
f_{\eta_c}^2 = 48 {m_c^2 \over M_{\eta_c}^3} |\Psi(0)|^2\; , \;\;\;\;\;\;\;
f_{J/\Psi}^2 = 12 {1 \over M_{J/\Psi}} |\Psi(0)|^2 \; ;
\end{equation}
therefore, the ratio $f_{\eta_c}/f_{J/\Psi}$ can be predicted in terms
of the meson masses and of the charm quark mass:
\begin{equation}
{f_{\eta_c} \over f_{J/\Psi}} =
2 m_c \Big( {M_{J/\Psi} \over M_{\eta_c}^3}\Big)^{1\over 2} = 0.97 \pm 0.03 \;;
\label{ratio}
\end{equation}
the deviations from the outcome of QCD sum rules,
at the level of $15- 20 \%$ for
$\eta_c$, $J/\Psi$ and $5 - 8 \%$ for
the radial excitations $\eta^\prime_c$, $\Psi^\prime$, can be
attributed to relativistic and radiative corrections to the constituent quark
model formula.
Tests of factorization can be performed by analyzing ratios of
decay widths, such as $B \to K^{(*)} \eta_c$ and
$B \to K^{(*)} \eta_c^\prime$, where the dependence on the Wilson
coefficients in the effective hamiltonian governing the decays,
and on other weak parameters drops out.
Let us consider, e.g., the ratio:
\begin{equation}
{\tilde R}_K = { \Gamma(B^- \to K^- \eta^\prime_c) \over
\Gamma(B^- \to K^- \eta_c)} = \; 0.771 \;
\big({f_{\eta^\prime_c} \over f_{\eta_c}}\big)^2 \;
\big({F_0(M^2_{\eta^\prime_c}) \over
F_0(M^2_{\eta_c})} \big)^2 = 0.60 \pm 0.15 \; .
\end{equation}
The interesting point is that, because of the flat shape of
$F_0(q^2)$ (fig.2a), ${\tilde R}_K$ mainly depends
on the ratio of the leptonic constants, so that
in factorization approximation a measurement of ${\tilde R}_K$
would provide us with interesting information on
${f_{\eta^\prime_c} \over f_{\eta_c}}$, and complement our knowledge of the
${c\bar c}$ wavefunction.
The analogous ratio for the decays into $K^*$ is given by
\begin{equation}
{\tilde R}_{K^*} = { \Gamma(B^- \to K^{*-} \eta^\prime_c) \over
\Gamma(B^- \to K^{*-} \eta_c)} =
0.381 \; \big({f_{\eta^\prime_c} \over f_{\eta_c}}\big)^2 \;
\big({A_0(M^2_{\eta^\prime_c})
\over A_0(M^2_{\eta_c})} \big)^2 =
0.381 \; \big({f_{\eta^\prime_c} \over f_{\eta_c}}\big)^2 \; (1.4 \pm 0.2)^2
\; .\label{polar}\end{equation}
\noindent
Here, the ratio of the form factors deviates from unity due to the
$q^2$-dependence of $A_0$ (fig.2b). The prediction from (\ref{polar})
would be:
${\tilde R}_{K^*} = 0.45 \pm 0.16$. Moreover,
the quantity $\sqrt{ {\tilde R}_{K^*}/{\tilde R}_{K}}$
is sensitive to the $q^2$-dependence of the ratio $A_0/F_0$:
\begin{equation}
1.42 \; \sqrt{ { {\tilde R}_{K^*}\over {\tilde R}_{K} } }=
\Big( { A_0(M^2_{\eta^\prime_c})/F_0(M^2_{\eta^\prime_c}) \over
A_0(M^2_{\eta_c}) /F_0(M^2_{\eta_c}) } \Big) \;.
\end{equation}
We also get:
\begin{equation}
{R}_{\eta_c} = { \Gamma(B^- \to K^{*-} \eta_c) \over
\Gamma(B^- \to K^- \eta_c)}= 0.373 \;
\big({A_0(M^2_{\eta_c})
\over F_0(M^2_{\eta_c})} \big)^2 =
0.73 \pm 0.13
\end{equation}
\noindent
and ${ R}_{\eta^\prime_c}=0.56 \pm 0.12$ for the analogous ratio
${ R}_{\eta^\prime_c}$.
Finally, the ratio:
\begin{equation}
R_K= { \Gamma(B^- \to K^{-} \eta_c) \over
\Gamma(B^- \to K^{-} J/\Psi)} =
2.519 \; \big({f_{\eta_c} \over f_{J/\Psi}}\big)^2 \;
\big({F_0(M^2_{\eta_c})
\over F_1(M^2_{J/\Psi})} \big)^2 \;
\end{equation}
\noindent can be predicted using
the simple pole model for $F_1^{B K}$.
We obtain:
$R_K=0.94 \pm 0.25$, and, for
$R^\prime_K= { \Gamma(B^- \to K^{-} \eta^\prime_c) \over
\Gamma(B^- \to K^{-} \Psi^\prime)}$:
$R^\prime_K= 1.61 \pm 0.53$.
Using the CLEOII measurements \cite{Hon}:
${\cal B}(B^- \to K^{-} J/\Psi) = (0.11 \pm 0.01 \pm 0.01)\times 10^{-2}$
and
${\cal B}(B^- \to K^{-} \Psi^\prime) = (0.06 \pm 0.02 \pm 0.01)\times 10^{-2}$
we expect:
${\cal B}(B^- \to K^{-} \eta_c) = (0.11 \pm 0.03)\times 10^{-2}$
and
${\cal B}(B^- \to K^{-} \eta_c^\prime) = (0.10 \pm 0.05)\times 10^{-2}$,
that should be within reach of present experimental facilities.
The measurement of some of the above decay rates
could shed more light on the problem of factorization,
which is a basic assumption in the present analysis of heavy meson nonleptonic
decays.
\noindent {\bf Acknowledgments. }
It is a pleasure to thank F.De~Fazio, C.A.Dominguez, G.Nardulli, N.Paver and
P.Santorelli for their collaboration on the subjects discussed here.
|
\section{Introduction}
Since the discovery
of high--$T_c$ superconductors\cite{bendorz}, the $S=1/2$ quantum
Heisenberg model has received considerable attention.
This is largely due to well-accepted experimental evidence
that
suggests that these compounds can be described\cite{manousakis} by a
two--dimensional (2D) doped Heisenberg
model
for spin $S=1/2$.
One of the experimental techniques used to study the excitations of
these systems is Raman scattering.
This technique was intensely used in the past for different
antiferromagnets of spin $S\ge 1$.
For such systems, the main features of the
line shape were explained by
Parkinson\cite{parkinson}, who used a spin--wave theory to
account for the photon--two-magnon process involved in the
Raman scattering. Very good agreement\cite{fleury} was found for
K$_2$NiF$_4$, which
is well described by a spin $S=1$ Heisenberg antiferromagnet.
The results of Raman experiments in the
parent insulating compounds of high--$T_c$ superconductors,
La$_2$CuO$_{4}$ and YBa$_2$Cu$_3$O$_{6.2}$, show some qualitative
differences with the line shape of previously studied
antiferromagnets\cite{sugai,sulewski,lyons}.
As in the case of K$_2$NiF$_4$, the line shape is centered at
an energy corresponding to a spin--pair excitation. However, in contrast
with the case of spin $S=1$, the linewidth is very broad, and the spectrum
has
a very long tail that extends beyond the energy of four magnon excitations.
Moreover, while the dominant contribution to scattering is in the so--called
$B_{1g}$ channel, there is also a significant contribution in
the nominally forbidden $A_{1g}$ configuration\cite{lyons}.
Since for lower spins the quantum fluctuations are larger,
some theoretical analysis beyond the mean--field spin--wave
approximation has been attempted
to explain the broad line shape.
Numerical diagonalizations in a $4 \times 4$ lattice shows very
little structure for the $B_{1g}$ channel: essentially three peaks;
a dominant two--magnon peak and two peaks identifiable as four--magnon
excitations\cite{GBC}. This calculation gives vanishing line shape for the
$A_{1g}$ channel.
Although the structure of the line shape is clearly different
from the one observed in the experiment, the first
three moments of both lines are in good agreement. The first three moments
of the distribution have been calculated in good agreement to experiment by
Singh {\it et al}.
using cumulant expansions on a Heisenberg model with diagonal
next--nearest--neighbor couplings\cite{singh}.
Canali and Girvin\cite{CG} used the Dyson--Maleev transformation
taking into account processes of up to four magnons, and presented convincing
evidence that the line shape cannot be explained by
quantum fluctuations alone.
Raman scattering Hamiltonians with a four--site
exchange\cite{honda} term have been proposed, these increase
the relative weight of the four--magnon
scattering, but does not explain the broadening of the peas.
This is also
consistent with
the work of Sugai\cite{sugai1}, and Roger and Delrieu\cite{roger}.
Also, recent studies of spin--pair excitations in a spin $S=1$ system,
NiPS$_3$, show a similar linewidth as those observed
in the cuprates\cite{merlin}.
From the above considerations, one concludes that it is necessary
to invoke an additional process. It was emphasized by
Nori {\it et al}.\cite{nori}
that the spin--phonon interaction
can be responsible for the broad linewidth, and for the finite cross--section
in the otherwise absent $A_{1g}$ channel.
They supported their arguments by computing the Raman cross section
in finite lattices, using a Heisenberg model with random {\it static}
exchange integrals, and averaging over configurations with equal
weight.
This calculation is in the spirit of the adiabatic or Born--Oppenheimer
approximation\cite{ziman}, that neglects the
fluctuations of the phonon field.
Adding phonons to the 2D $S=1/2$ Heisenberg Hamiltonian (HAF)
provides a mechanism\cite{nori} for the otherwise
forbidden $A_{1g}$
scattering, as well as allowing a coupling between the ground and excited
spin states.
In a typical HAF, the
phonon frequency\cite{nori}
is about a third of
the interaction constant $J$. The first excited state lies
at $3J$, so there is an order of magnitude energy
difference between spin and phonon excitations.
It was argued\cite{nori} that the separation of energy scales justifies the
adiabatic phonon approximation.
In the present work we consider the Raman scattering for a Heisenberg
model with spin--phonon interaction, including the effects of the phonon
dynamics. We solve for the exact ground state of a system in which
the vibrational degrees of freedom are allowed a finite number
of excitations, an approximation valid for small but finite
phonon frequencies.
To the best of our knowledge, this is the first calculation
of exact diagonalizations for a dynamic spin--phonon system.
Some work including phonon dynamics beyond the adiabatic
approximation exists in
the context of one--dimensional Peierls systems. Fradkin and
Hirsh\cite{fradkin} studied the electron--phonon interaction
using quantum Monte Carlo
simulations, and Proetto and Falicov\cite{proetto}
solved the case of two sites and one phonon.
This paper is organized
as follows. In Sec. II we present our Hamiltonian and scattering operator
and discuss the rationale for our truncated--phonon model.
In Sec. III we present computational results and provide theoretical support
for these results.
We also present an
alternative approach to including phonons and compare it to our model.
Sec. IV is devoted to conclusions.
\section{Model and Procedure}
We study a Heisenberg model with spin--phonon interaction described by
the following Hamiltonian:
\begin{equation}
H = \sum_{<ij>} \left\{ \left[J - \alpha (a^{\dagger}_{ij}+a_{ij})\right]
\vec S_{i} \cdot \vec S_{j}
+ \omega_0 a^{\dagger}_{ij}a_{ij} \right \}
\label{H} \,
\end{equation}
where $\vec S_{i}$ are spin $1/2$ operators, $a^{\dagger}_{ij}
(a_{ij})$
is a creation (annihilation)
operator for an Einstein phonon of frequency $\omega_0$, and the simbol
$``ij"$
refers to a link of a square lattice.
In $J_{ij} = J - \delta J_{ij}$ we include the coupling
between the phonons and the spin degrees of freedom
through a ``displacement" operator
$\hat x_{ij} = \sqrt{\frac{\hbar}{2m\omega_0}}
(a^{\dagger}_{ij}+a_{ij})$.
Our
parameter $\alpha $ in
Eq.(\ref{H}) is proportional to the spin--phonon coupling constant
$\lambda=\alpha \sqrt{{2m\omega_0}/{\hbar}}$ relating the change in
the exchange integral $\delta J_{ij}$
with the displacement: $\delta J_{ij} =\lambda \hat x_{ij} = \alpha
( a_{ij}^{\dagger} + a_{ij})$.
Due to computational limitations, we need to restrict ourselves to a small
number ($6$ and $8$)
of spins. To have these as two dimensional as possible, we
placed them in
non-periodic ladder type structures as shown in Fig. \ref{Figure1}.
We use Einstein phonons to simplify the model, and we consider the
highest--energy
phonons as they are the closest in energy to the magnon excitations
and will thus have the greatest coupling to the magnons.
The occupation number of the phonon degrees of freedom at each
link is in principle unrestricted; the number of phonons ranges
from zero to infinity. This makes the problem intractable from the point
of view of exact diagonalizations,
since the resulting Hilbert space is infinite.
In order to overcome this limitation, we restrict the possible
number of phonons at each link
by imposing the
condition $(a^{\dagger})^{n} = 0$. Computational resources limit us to
using $n = 2,3$.
This approximation maps the phonon degree of freedom into a two-- or
three--state system
at each link.
The truncation of the phononic occupation states implies that the
variations of $J_{ij}$ due to quantum fluctuations
are bounded. For example, for $n=2$,
the maximum displacements for a given link is given by the ``coherent"
states
$|\psi_{ij}^{\pm}\rangle=2^{-1/2}(1\pm a^{\dagger}_{ij})|0\rangle$,
and $\delta J_{ij}= \alpha \langle \psi_{ij}^{\pm}|
(a^{\dagger}_{ij}+a_{ij})|\psi_{ij}^{\pm}\rangle=\pm\alpha$.
Quantum fluctuations themselves are certainly limited by
our
truncation, but since they are in general small for small $\omega_0$,
our approximation will account for the relevant dynamics in the
regime $\delta J_{ij} /J < 1$ and $\omega_0 /J <1$.
We are interested in the Raman scattering intensity $
I_R(\omega)$, given by:
\begin{equation}
I _R(\omega)=\sum_{\nu}\: |\langle \nu|\hat R|0\rangle|^{2}
\:\delta(\omega-
E_\nu+E_{o})
\label{sigmar} \, ,
\end{equation}
where $|\nu\rangle$ are the eigenstates and $E_\nu $ the eigenvalues
of $H$.
We
compute $ I _R(\omega)$
by using the partial
fraction expansion method of Gagliano and Balseiro\cite{GB}.
The relevant operator $\hat R$ in Eq.(\ref{sigmar})
depends on the different configurations of the Raman
experiment. In general\cite{parkinson}:
\begin {equation}
\hat R = \sum_{<ij>} (\vec E_{\rm inc}\cdot\hat x_{ij})
(\vec E_{\rm scat}\cdot\hat x_{ij})
\vec S_{i}\cdot \vec S_{j}
\label{A} \, ,
\end{equation}
with $\vec E_{\rm inc}$ and $\vec E_{\rm scat}$ refer to the polarization
of
the incoming and scattered photon respectively, and $\hat x_{ij}$
are unit vectors in the near--neighbor directions of the lattice.
For square lattices, two common
scattering configurations
are the so--called $B_{1g}$ and $A_{1g}$ symmetries.
We let the lattice axis lie along the $x$
and $y$ directions, and we define $x'$ and $y' $ the directions along
the lattice diagonals ($\hat{x'} = \frac{\hat{y}+\hat{x}}{\
\sqrt{2}}$
and $\hat{y'} = \frac{\hat{y}-\hat{x}}{\sqrt{2}}$),
then $B_{1g}$ corresponds to
$\vec E_{\rm scat} \parallel \hat{x'}$ and $\vec E_{\rm inc} \parallel
\hat{y'}$,
whereas $A_{1g}$ corresponds to both the incident and scattered photon
polarized
in the same direction: $\vec E_{inc}\parallel \vec E_{\rm scat} \parallel
\hat{x'}$.
For the case of the ``pure" Heisenberg Hamiltonian
(no phonons, or
$\alpha =\omega_0=0$),
the
absorption
corresponding to the $A_{1g}$ symmetry is zero at any non-zero frequency,
since the $A_{1g}$ operator is proportional to the Hamiltonian.
We show in the next section how the inclusion of the phonon dynamics
gives rise to a finite absorption in this channel, and
argue that a spin--phonon interaction accounts for the
main features observed in the Raman experiments in the undoped
copper oxides.
Also, we compare the results obtained in the dynamical model
with that of the Heisenberg model with random (static)
exchange integrals, which was studied in this context by Nori
{\it et al}.\cite{nori}
\section {results and discussion}
In this section we present
numerical results for systems of $6$ and $8$ spins\cite{hilb}
corresponding to the geometry shown in Figure \ref{Figure1}.
We first discuss the $A_{1g}$ symmetry. In this case, and in
the absence of lattice distortions, the Raman operator
commutes with the Heisenberg Hamiltonian.
No line shape is observed in this case:
$I_R(\omega) \sim \delta (\omega)$.
It was first pointed out by Nori {\it et al}.\cite{nori}
that the presence of {\it static}
disorder in the exchange integrals $J_{ij}$ changes the value of the
commutator, and
can give rise to a finite $A_{1g}$ signal quite similar to experiments.
It was argued by Nori {\it et al}.\cite{nori} that the spin--phonon coupling
produces disorder in the values of $J_{ij}$,
in the limit where the vibrational
motions are much slower than the spin degrees of
freedom. Here, we are interested in how this limit is achieved
in a system that---up to boundary
effects---is translationally invariant, and includes the
{\it dynamics} of the spin--phonon coupling.
Consider our Hamiltonian Eq.(\ref{H}) and the operator $\hat R$ for
the $A_{1g}$ symmetry.
For $\omega_0 \neq 0$, $[H,\hat R] \neq 0 $ and we expect a finite
line shape. In Fig. \ref{Figure2} we show the $I(\omega)$ obtained
for $\alpha=0.1J$ and $\omega_0=0.05J$.
Some of the qualitative features of the experimental line shape are already
present in these finite--size systems: there is a broad
line shape of width $\sim 2J$ with a maximum at the two--magnon
energy. Note that for our ladder geometries,
since most sites have coordination number $3$, the two--magnon
energy in the Ising limit is located at $2J$, instead
of the corresponding $3J$ of the square lattice.
Due to the finite size of the lattice, the line shape consists
of a series of peaks that are centered at the unperturbed
energies, which are indicated by arrows in Fig. \ref{Figure2}. These
unperturbed energies
correspond to the manifold of two--and multi--magnon states.
In turn, these ``internal" peaks have a finite width that
is due to the coupling with the phonons.
The energy scale dominating this width is $\alpha$.
In order to test the validity of the Hilbert space truncation,
we have increased the number of allowed excitations per phonon by one, with
no qualitative changes in the line shape (see Fig. \ref{Figure3})
It is interesting to consider the following two
limiting cases: $(a)$ vanishing spin-phonon coupling $\alpha \rightarrow 0$
for finite phonon frequency $\omega_0$,
and $(b)$ vanishingi
phonon frequency $\omega_0 \rightarrow 0$ for finite $\alpha$.
In both limits the $A_{1g}$ line shape vanishes.
In $(a)$, the spins are uncoupled from the phonons,
and the line shape vanishes because in this
limit $[H,\hat R]=0$.
Figure \ref{Figure4} illustrates how the line shape
vanishes and shifts to lower energies as $\alpha \rightarrow 0$.
In limit $(b)$, the phonons are ``static", but remain coupled
to the spins, because $\alpha \neq 0$. Since the phonons have
no dynamics, the eigenstates can be written in the form
\begin{equation}
|\psi _\nu (\omega _0=0)\rangle =
|\psi _\nu \{n\}\rangle \otimes |\{n\}\rangle \, ,
\end{equation}
where $|\{n\}\rangle$ is a static configuration of displacements,
in such a way that, for each link $ij$, one has
$(a^{\dagger}_{ij}+a_{ij}) |\{n\}\rangle=
\eta_{ij}({\{n\}}) |\{n\}\rangle$, with $\eta_{ij}(\{n\})$ a $c$--number.
The state $|\psi _\nu \{n\}\rangle$ involves spin degrees of freedom
only, and is an eigenstate of the following Hamiltonian
\begin{equation}
H\{n\} = \sum_{<ij>}
\left[J -\alpha \eta_{ij}\left(\left\{n\right\}\right)\right]
\vec S_{i} \cdot \vec S_{j}
\, .
\end{equation}
For $\omega_0=0$, the
problem is that of an {\it annealed} configuration of displacements:
for each configuration $\{n\}$, the couplings $J_{ij}$ are different,
and one has to solve for the spin dynamics in the presence of
this distribution of couplings. The complete spectrum is obtained
by solving $N_p$
decoupled eigenvalue problems, with $N_p$ being
the phononic Hilbert space dimension. For our case of one
excitation per link, $N_p = 2^{N_{\rm links}}$, with
${N_{\rm links}}$ being the number of links of the lattice.
What distinguishes this from a quenched disorder is the fact that
the ground state of this problem corresponds to the lowest energy
state of the spin system in an {\it ordered} background of
couplings. Since this is also an eigenstate of $\hat R$ for
the $A_{1g}$ symmetry, the line shape will be zero. In Fig. \ref{Figure5}
we show how the line shape vanishes as $\omega_0$ decreases to
zero.
Note that from this analysis one concludes that,
even in the case of $\omega_0=0$, the line shape
is non--zero at finite temperature. This is due to the presence of
thermally excited states, which are eigenstates of a Hamiltonian
that does not commute with $\hat R$.
The right column
of Figure \ref{Figure5} shows the rescaled results, which illustrate
the line shape approaching a limiting function of $\omega$ whose
overall amplitude
vanishes as $\omega_0 \rightarrow 0$. The main features of this
limiting function
should be accounted for by a model of static disordered couplings.
In previous work in the context of one--dimensional Peierls
systems\cite{wilkins_igor}, it is argued that
in the limit of small phonon frequencies,
the quantum lattice fluctuations can be modeled by a static,
random Gaussian potential with zero mean.
A model in which the exchange integrals are taken from a random configuration
of couplings was also studied by Nori {\it et al}\cite{nori}.
We can test this hypothesis in
our dynamical
model. Results of the comparison between the dynamical and a static
disordered system are shown in Fig. \ref{Figure6}.
Note that
both the amplitude
and the position of the peaks are in very good mutual agreement.
The scattering peaks from the two models have their energies scaled with
respect
to each other. After they are rescaled, the
location and magnitude of
the peaks are very close for both models (see Fig. \ref{Figure6}).
The basis for this rescaling of the horizontal axis is the following.
For our
dynamic phonon model, the
minimum
energy state corresponds to the lattice ``maximally compressed":
$\eta_{ij}=-\langle x \rangle_0$.
The tendency of the system to compress has the
effect of
renormalizing the
interaction coefficient $J$, since the displacement $x_{
ij}$
will tend to
be a constant, and therefore $J' = J + \alpha \langle x \rangle_0$. The
average displacement will be
greater
as the possible number of excitations increases; this
explains a shift in
energy between $1$ and $2$ excitations as well.
In the model with
static disorder, the
phonon energy goes to zero, but the fluctuations are still present.
In the treatment of Nori {\it et al}.\cite{nori}
the argument
used is that the phonon energy is small compared to that of the spin
excitations,
and thus can be neglected.
However,
we have shown that for $\omega _0 $ strictly zero the line shape vanishes.
At zero temperature,
the disordered model should be compared with the dynamical model
at infinitesimal $\omega_0$. We prove this statement
by
using perturbation theory in $\omega_0$. Assume an ordering of the states
for $\alpha =\omega_0 =0$, and let us label those states by the index
$\nu$. If $\alpha \ll J$ and $\omega_0 =0$,
each state $|\psi_\nu \rangle $ splits in a manifold
$|\psi_\nu\{n\} \rangle $ of almost degenerate states of energies
$E_\nu\{n\}$, that correspond to the ``disordered" configurations.
If we now turn on $\omega_0$, different states are going to mix,
in such a way that the ground state can be written as
\begin{equation} |\psi_0\rangle \simeq |\psi_0\{n\}_0 \rangle +
\omega_0\sum_{\{n\}\neq\{n\}_0} c(\{n\})|\psi_0\{n\}\rangle
\otimes |\{n\}\rangle \, ,
\end{equation}
where $\{n\}_0$ is the ordered configuration of couplings.
A similar expansion can be used for the excited states.
If we use the fact that $A_{1g}$ $\hat R$ acts only on the spin degrees of
freedom, and that different states $ |\{n\}\rangle$ are orthogonal,
to lowest order in $\omega_0$, $I(\omega)$ will be given by
\begin{equation}
I(\omega)\simeq \omega_0^2\sum _{\{n\}} |c(\{n\})|^2
\sum_\nu |\langle \psi_\nu\{n\}|\hat R |\psi_0\{n\}\rangle|^2
\delta(\omega -E_\nu\{n\}+E_0\{n\}) \, .
\end{equation}
We have found
computationally a falloff of $I(\omega)$ proportional to $\omega_0^2$.
Now, if we keep the assumption of $\alpha \ll J$, the states corresponding
to each manifold are almost degenerate and we can approximate
$c(\{n\}) \sim (M-1)^{-1/2}\sim {M}^{-1/2}$,
with $M$ the
number of configurations $\{n\}$. Therefore, in this limit $I(\omega)$
is given by an average over configurations with {\it equal} weight.
We now
present results
for the $B_{1g}$ configuration. In Fig. \ref{Figure7} we compare
the
$B_{1g}$ scattering from eight $S=1/2$ Heisenberg spins
for three different cases: bare HAF, with phonon interactions,
and with static disorder.
We have argued that peaks at higher excitations will eventually merge
into one broad peak
as the lattice
size increases. To show that this is the case, we have calculated
the $A_{1g}$ and
$B_{1g}$ peaks
for a $3\times 4$ non--periodic lattice with static disorder. The
results are shown in Fig. \ref{Figure8}.
These are in good qualitative agreement
with experiments,
except that the experimental scattering is centered at $\omega \sim 3J$
instead of $2J$ in our case.
This is due to finite size effects and to the lattice being non--periodic.
The model with static
disorder, when extended to larger numbers of spins than can be obtained
with the
dynamic phonon model, begins to duplicate experiments. The agreement
between the static and
dynamic phonon
models suggests that the dynamic phonon model will also agree with
experiments.
\section{Conclusions}
In this paper we have presented a model of phonon-magnon interaction with
a truncated phonon space.
We discussed how this model can explain the otherwise
forbidden $A_{1g}$
scattering and showed that it does in fact give rise to $A_{1g}$
scattering. We discussed how coupling to spin excitations leads
to a broadening
of the line shape. We have proven that phonons can be modeled by
static disorder and compared
the results obtained by using quenched phonons to our dynamical model.
We showed that the amplitude of the $A_{1g}$ scattering should fall off
as $\omega_0^2$.
We showed that both the $A_{1g}$ and $B_{1g}$ scattering is broadened.
In conclusion,
the broadening is due
to coupling to spin excitation states through the phonon interaction.
\section{acknowledgements}
We would like to
thank Roberto Merlin and Jim Allen for helpful discussions, and
Franco Nori for helpful discussions and a careful reading of this manuscript.
|
\section{ INTRODUCTION}
The investigation of the symmetry of the superconducting
order parameters
is currently one of the most exciting problems in the
field of
high-$T_c$
research.~\cite{MBK92,SDW93,HBM93,KCN94,DES94,WVL93,CSY94,BO94,TKC94}
Many experiments suggest that the order
parameter is
anisotropic.~\cite{MBK92,SDW93,HBM93,KCN94,DES94}
Furthermore, the observation of $\pi$ phase shifts
in corner junctions
\cite{WVL93,BO94,TKC94}is consistent with a
$d_{x^2-y^2}$
symmetry of the gap function, although another
experiment~\cite{CSY94}
favors a more conventional $s$-wave pairing.
On the other hand, the
occurrence of a finite tunneling current perpendicular
to the planes
\cite{SGM94} seems to be incompatible with $d$-wave
symmetry in a simple
picture.
In view of this debate, it is important to develop
alternative
experimental techniques which are able to discriminate
between the
various symmetries of the gap function.
In particular, it is interesting to measure the
symmetry of the gap function
without the necessity of a tunneling contact, which
always
includes the problems of a residual magnetic field,
trapped flux in the
tunneling circuit, or singularities in the supercurrent
flow at the
corner~\cite{VH95}.
On the other hand, optical second harmonic generation
has widely been used
as a probe for two dimensional optical, electronic and
structural properties at various interfaces such as
solid-vacuum,
solid-gas, solid-solid, solid-liquid, or
liquid-gas~\cite{shen,richmond}.
Recently, the
nonlinear magneto-optical Kerr-effect has become a
new nonlinear optical
method for the investigation of low-dimensional magnetic
structures~\cite{pan,hub2,hub3}.
This effect is interface sensitive within the electric
dipole-approximation
and directly probes, unlike methods based on Raman
scattering, low-energy
magnetic excitations ($\sim$ meV) using optical
interband transitions
($\sim$ eV). In contrast to linear optical techniques,
the nonlinear
magneto-optical Kerr effect yields an excellent signal
to noise ratio,
since it is free of linear nonmagnetic background
radiation. This
advantage manifests itself both in the large magnetic
intensity
contrast upon magnetization inversion,
\[\frac{I({\bf M})-I(-{\bf M})}{I({\bf M})+I(-{\bf M})}
\]
of about 40 \% for Fe surfaces\cite{reif,hub4} compared
to a linear effect of
typically 0.1 \% and in a giant nonlinear Kerr rotation
$\phi_{K}^{(2)}$ which,
depending on the chosen Kerr-configuration, can be
tuned up to 90$^{\circ}$
nearly at will~\cite{pusto,kirschner,theo}.
This corresponds to an enhancement by 2 - 3 orders of
magnitude compared to
the usual linear Kerr rotation, which is even reduced
in thin films.
Thus, it allows for the determination of the magnetic
interface symmetry,
including the magnetic ``easy axis''~\cite{hub5}.
It is well-known that the symmetry of the nonlinear
optical susceptibility is
strongly affected by a magnetization or an external
applied magnetic
field.~\cite{pan} Thus it is of considerable interest
to extend the symmetry
analysis of the nonlinear magneto-optical response to
the superconducting state
for different symmetries of the superconducting order
parameter.
Therefore, we propose in this paper a new theory for
the nonlinear
optical response of a superconductor in the presence
of a magnetic field,
which is able to distinguish
between certain symmetries of the gap function and
might stimulate
corresponding experiments.
In the following, we present our theory and show that
indeed
a symmetry dependent contribution to the magneto-optical
response without
tunneling contact results in optical second harmonic
generation.
Although it is, due to the gauge invariance,
impossible to measure
the phase of the superconducting order parameter
without tunneling
contact, it is still possible to measure its
symmetry, not solely
its magnitude.
This results from the interference of different pairing
amplitudes in the dipole
matrix elements of the {\em three} transitions in
nonlinear optics.
Due to the surface sensitivity of nonlinear optics
for systems with bulk
inversion symmetry, one can take advantage of the
broken inversion symmetry at the surface.
This is of interest, because Cooper pairing,
together with the
always present spin-orbit coupling, is then no
more purely
singlet- or triplet-like.
The interference of the singlet and triplet pairing
states, which is linear in
spin-orbit interaction, leads to the symmetry sensitive
contribution of the nonlinear optical response for
systems in an external magnetic field.
Note that the possible importance of mixed singlet
and triplet pairing states at
the interface of a tunneling contact between a
heavy fermion and a conventional
superconductor has already been discussed by
Fenton\cite{F85}.
In order to obtain a more systematic insight
into these phenomena,
we perform the description of a superconductor at
the surface of a bulk inversion symmetric system,
including spin-orbit
interaction and give a group-theoretical classification
of the irreducible
representations of the gap function.
This is of importance for a detailed calculation of
the tensor elements of the
nonlinear optical susceptibility which is performed
in the second part
of the theory.
Finally, we present our results concerning the
symmetry dependence of the
corresponding experiment and discuss in detail
the differences
of the line shapes of the optical spectra.
We find that it is possible to discriminate between
an isotropic $s$-wave,
a $s_{x ^2 + y ^2}$- or $d_{x ^2 - y ^2}$- wave
and a $d_{x y }$- wave.
However, no symmetry dependent differences between
the $s_{x ^2 + y ^2}$- and
the $d_{x ^2 - y ^2}$- waves occur, which are
actually the most discussed
symmetries of the high-T$_c$ systems.
Nevertheless nonlinear magneto-optics is shown
to be an
alternative and complementary method to gain
insight into the
symmetry of the superconducting gap function.
\section{ THEORY}
The strategy of this section is to calculate the
effects of $s$- and $d$-wave
superconductivity on the nonlinear optical
susceptibility tensor $\chi^{(2)}$,
which are due to (a) the modification of the
bandstructure and thus the
resonance denominators and (b) due to the
symmetry of the superconducting
order-parameter affecting the wave functions
of the optical electric-dipole
transition matrix elements. Thus, we proceed
as follows: (i) First we set up
the superconducting BCS-type Hamiltonian and
perform its group theoretical
analysis. In order to obtain the desired
sensitivity of nonlinear optics
to the symmetry of the gap function this
requires, as it will turn out,
the simultaneous absence of inversion symmetry,
presence of an
external magnetic field breaking time reversal,
and spin-orbit interaction
coupling singlet and triplet pairs. (ii) Making
use of this symmetry
classification and the mentioned constraints we
then calculate the nonlinear
magneto-optical response in $s$- and $d$- wave
superconductors from the
appropriate current-current-current correlation
function and propose
a suitable experimental geometry for the observation
of this new nonlinear
magneto-optical effect in superconductors.
In our theory for the nonlinear magneto-optical
response of unconventional
superconductors, the superconducting state is
described within the BCS
theory~\cite{BCS57}. The corresponding Hamiltonian
with arbitrary pairing
symmetry and in a magnetic field $\vec{h}$ is
given by
\begin{equation}
H=\sum_{{\bf k}\mu} \Psi_{{\bf k}\mu}^\dagger
{\cal H}_{{\bf k}\mu}
\Psi_{{\bf k}\mu}\, ,
\end{equation}
with the four-component Nambu spinor
$\Psi_{{\bf k} \mu} = (c_{{\bf k} \mu \uparrow},
c_{{\bf k} \mu \downarrow},
c^\dagger_{-{\bf k} \mu \uparrow},
c^\dagger_{-{\bf k} \mu \downarrow})$.
Here, $c^\dagger_{{\bf k} \mu \sigma}$ is the
creation operator of an electron
with momentum ${\bf k}$, band index $\mu$ and
spin $\sigma$.
The $(4 \times 4)$ matrix ${\cal H}_{{\bf k}\mu}$
can be expressed in terms
of $(2 \times 2)$ block matrices:
\begin{equation}
{\cal H}_{ {\bf k} \mu }= \left( \begin{array}{cc}
\varepsilon_{{\bf k} \mu}
\hat{\sigma}^o - \vec{h} \cdot
\hat{\vec{\sigma}} &
\hat{\Delta}_{{\bf k}\mu} \\
-\hat{\Delta}^\ast_{-{\bf k}\mu} &\
-\varepsilon_{{\bf k}\mu}
\hat{\sigma}^o +\vec{h}\cdot
\hat{\vec{\sigma}}^\ast
\end{array} \right) \, .
\label{HBCS}
\end{equation}
The block matrices are expanded in terms of
the unit matrix $\hat{\sigma}^o$
and the vector of the Pauli matrices
$\hat{\vec{\sigma}}$.
This notation is close to that of Sigrist
and Ueda~\cite{SU91}.
The symmetry of the superconducting order
parameter is characterized by the
gap function
$\Delta_{\sigma \sigma ' ;{\bf k}\mu}=
\langle c_{{\bf k} \sigma \mu}
c_{-{\bf k} \sigma ' \mu } \rangle$.
We neglect any diamagnetic, i.e. Meissner
effect of the magnetic field,
but assume a large penetration depth at the
surface and no influence of
the vortex structure to the optical spectrum.
This seems to be reasonable at least for the
excitations in the interband
regime discussed in this paper.
$\Delta_{\sigma \sigma' ;{\bf k}\mu}$ is
decomposed in the usual way in singlet states
($\Delta ^o_{{\bf k} \mu}=\Delta ^o_{-{\bf k}\mu}$)
and triplet states
($\vec{d}_{{\bf k}\mu}=-\vec{d}_{-{\bf k}\mu}$):
\begin{equation}
\hat{\Delta}_{{\bf k}\mu}=\left(\Delta ^o_{{\bf k}\mu}
\hat{\sigma}^o +
\vec{d}_{{\bf k}\mu} \cdot \hat{\vec{\sigma}} \right)
i\hat{\sigma} ^y \, .
\label{gapexp}
\end{equation}
The symmetry of $\Delta ^o_{{\bf k}\mu}$ and
$\vec{d}_{{\bf k}\mu}$
with respect to the
transition from ${\bf k}$ to $-{\bf k}$ is a
direct consequence of the
Pauli principle.
Since we consider the states at a surface, ${\bf k}$
refers to the
two dimensional in-plane momentum.
Below the transition temperature $T_c$, the symmetry
of a system is reduced
compared to the high temperature phase.
The symmetry group $G$ of the high temperature phase
is determined by the
symmetry operations which keep the Hamiltonian for
$ \hat{\Delta}_{\bf k}=0$
invariant.
We consider a system which is for $\vec{h}=0$
invariant with respect to the group
\begin{equation}
G=g \times K \times U(1) \, ,
\end{equation}
where $g$, $K$ and $U(1)$ are the point group,
time reversal operation and the
gauge group of multiplication of electron creation
operator by an arbitrary
phase, respectively.~\cite{SU91,A85,VG85,B85,SR89,SR92}.
In the ordinary case the normal-state gauge
symmetry is broken
at the superconducting phase transition, i.e.
the residual symmetry group is
$g \times K$. This is called a conventional
superconductor.
In unconventional superconductors however,
the symmetry is lower than
$g \times K$. At the transition temperature,
the BCS - gap equation is an
eigenvalue equation and consequently, an eigenvector
$\Delta_{\sigma \sigma' ;{\bf k} \mu }$ belongs to
one of the irreducible representations
${\cal D}$ of the group $G$.
If ${\cal D}$ is the unit representation
${\cal A}_1$ (or ${\cal A}_{1g}$
for systems with inversion symmetry) conventional
superconductivity occurs.
In all other cases the superconductivity is
unconventional.
In order to discuss the various symmetry states
of the order parameter,
one has to generate all irreducible representations
of the gap function,
where, due to spin-orbit coupling, the spin
degrees of freedom cannot be
transformed independently from the spatial
(orbital) coordinates.
For various point groups this symmetry
classification has been
performed~\cite{SU91,A85,VG85,B85,SR89,SR92}.
In all these cases the inversion
operation $C_{i}$ is an element of the group $G$.
Since we are interested in the investigation of
superconducting properties with
surface sensitive nonlinear optical experiments,
we have to take the effect of
broken inversion symmetry into account.
In order to be specific, we consider the surface
of a tetragonal system
(bulk point group D$_{4 h}$) with residual point
group $C_{4v}$.
This group has five irreducible representations:
four of dimension one
(${\cal A}_1$, ${\cal A}_2$, ${\cal B}_1$ and
${\cal B}_2$) and one of
dimension two (${\cal E}$).
The isotropic $s_o$ and and anisotropic
$s_{x^2+y^2}\; $s-waves transform
as ${\cal A}_1$, the $d_{x^2-y^2}$-wave as
${\cal B}_1$,
the $d_{xy}$-wave as ${\cal B}_2$, a
$d_{x^2-y^2}d_{xy}$-wave
as ${\cal A}_2$ and the $p_x$, $p_y$ waves
as ${\cal E}$.
In order to classify the irreducible
representations of the gap function, we
have to analyze the transformation properties of
$\Delta_{\sigma \sigma' ;{\bf k}\mu}$.
Applying an element $R$ of the point group to
$\hat{\Delta}_{{\bf k}\mu}$,
the following transformation of the singlet and
triplet part results:
\begin{equation}
R\, \hat{\Delta}_{{\bf k}\mu}= \left(
\Delta ^o_{{\cal D}^{(1)}_R{\bf k} \mu} +
\tilde{{\cal D}}^{(1)}_R
\vec{d}_{{\cal D}^{(1)}_R{\bf k}\mu}
\cdot \hat{\vec{\sigma}} \right) i\hat{\sigma ^y} \, .
\end{equation}
Here, ${\cal D}^{(1)}_R$ is the representation
of $R \in G$ which transforms the
coordinates.
If one considers the transformation of a Pauli
spinor with respect to a
combination $R=R_o C_i$ of the inversion
operation $C_i$ and a rotation
$R_o$, only the rotational part has to be
applied to the spinor, i.e. the
representation of $R$ in spin space is
${\cal D}^{(1/2)}_{R_o}$.
Consequently, for the vector in spin space
$\vec{d}_{\bf k}$, the representation
$\tilde{{\cal D}}^{(1)}_R \equiv {\cal D}^{(1)}_{R_o}$,
where the inversion
operation is replaced by the identical
transformation, has to be applied.
Therefore, one finds in the {\em bulk} of a
system with inversion symmetry:
$ C_i\, \hat{\Delta}_{{\bf k}\mu}=
\left(\Delta ^o_{{\bf k}\mu} -
\vec{d}_{{\bf k}\mu}
\cdot \hat{\vec{\sigma}} \right) i\hat{\sigma ^y}$,
since the vector $\vec{d}$ is not affected
directly by the inversion operation.
The minus sign results from the inversion
of ${\bf k}$ to $-{\bf k}$.
Consequently, in the {\em bulk} , the singlet
and triplet part belong to
different irreducible representations and
either singlet or triplet
superconductivity occurs.
In distinction to this, a coexistence of singlet
and triplet pairing states is
possible for systems without inversion symmetry,
i.e. at the {\em surface}.
Now, the in-plane inversion operation can
be realized by a rotation
(rotation by $\pi$ with $z$-axis as rotation axis).
This rotation transforms the vector $\vec{d}$
to $-\vec{d}$ and
the minus sign of the transformation
${\bf k} \rightarrow -{\bf k}$
is canceled. Consequently, the irreducible
representations of the gap function
at the {\em surface} contain both, singlet
and triplet parts.
{}From these considerations one obtains the
irreducible representations of
the gap function from the simultaneous
Clebsch-Gordon coupling of orbital
and spin degrees of freedom.
The results for the simultaneously occurring
singlet and triplet
part of the pairing amplitude are given in
the table.
Sigrist and Rice~\cite{SR92} calculated the
irreducible
representations of the tetragonal group
$D_{4h}$. Since
$D_{4h}=C_{4v} \times C_i$, it is
straightforward to check the above results
by reducing the subduced representations of $D_{4h}$.
One finds: ${\cal A}_1={\cal A}_{1g} \oplus {\cal A}_{2u}$,
${\cal A}_2={\cal A}_{2g} \oplus {\cal A}_{1u}$,
${\cal B}_1={\cal B}_{1g} \oplus {\cal B}_{2u}$,
${\cal B}_2={\cal B}_{2g} \oplus {\cal B}_{1u}$ and
${\cal E}={\cal E}_{g} \oplus {\cal E}_{u}$ which
leads to the results of the
table. Here, $g$ and $u$ refer to the irreducible
representations of $D_{4h}$
with even and odd parity.
The irreducible representations of table~\ref{tab1}
are the possible symmetry
states of a superconductor on the surface of a
tetragonal system and with spin
orbit interaction. If one neglects the spin orbit
coupling, the spin and orbital
degrees of freedom transform separately and the
singlet and triplet pairing
states decouple again. Therefore, for a bulk
singlet superconductor, the
simultaneously occurring triplet part at the
surface is of the
order of the spin-orbit interaction, and vice
versa.
In heavy fermion systems the spin-orbit interaction
is large, but even in
transition metals one expects this quantity to
be of the order of 50 meV which,
although being smaller, has nevertheless dramatic
consequences such as the
reorientation of the magnetic easy axis in thin
ferromagnetic films upon
the increase of the film thickness or the rise
of the temperature.
Finally, we expect the spin-orbit induced triplet
part to be observable if one
considers a surface sensitive experiment such as
second harmonic generation.
Based on these group theoretical classifications,
we calculate now
the nonlinear magneto-optical susceptibility tensor
of a superconductor
and focus on the interference of the simultaneously
occurring singlet and
triplet part of the gap function.
The optical response in second harmonic generation can
be obtained from the
nonlinear current-current-current correlation function:
\begin{eqnarray}
\chi_{\alpha \beta \gamma}({\bf q}, \omega) &=&
\int_{-\infty}^{\infty} \frac{d \epsilon}{\pi}
\int_{-\infty}^{\infty} \frac{d \epsilon '}{\pi}
\int_{-\infty}^{\infty} \frac{d \epsilon ''}{\pi}
I_{\alpha \beta \gamma}({\bf q},\epsilon,\epsilon ',
\epsilon '') \;\times
\nonumber \\
& &
\frac{
\frac{f(\epsilon '')-f(\epsilon ')}
{\omega +i\delta -\epsilon '' +\epsilon '} \, - \,
\frac{f(\epsilon' )-f(\epsilon )}
{\omega +i\delta -\epsilon '
+\epsilon}}
{2(\omega +i\delta) -\epsilon '' +\epsilon }
\end{eqnarray}
where $f(\epsilon)$ is the Fermi function and
the spectral
function
$I_{\alpha \beta \gamma}
({\bf q},\epsilon,\epsilon ',\epsilon '') $
is given by
\begin{equation}
I_{\alpha \beta \gamma}
({\bf q},\epsilon,\epsilon ',\epsilon '')
={\rm Tr} \left( J_{-2{\bf q} \alpha}
\varrho(\epsilon)
J_{{\bf q}\beta} \varrho(\epsilon')
J_{{\bf q} \gamma} \varrho(\epsilon'')\right) \, .
\label{nolispec}
\end{equation}
$J_{{\bf q} \alpha}$ is the $\alpha$-th
component of the
current operator
\begin{equation}
\vec{J}_{\bf q}= \sum_{{\bf k} \sigma \mu \nu}
\vec{j}_{{\bf k}\mu \nu}
c^\dagger_{{\bf k}+\frac{{\bf q}}{2} \sigma \mu}
c_{{\bf k}-\frac{{\bf q}}{2} \sigma \nu}\, ,
\end{equation}
and $\varrho(\epsilon)=-\frac{1}{\pi}{\rm Im}
(\epsilon+i\delta +H)^{-1}$
is the density of states matrix with
Hamiltonian $H$ of Eq.~\ref{HBCS}.
The trace has to be performed with respect
to all single particle states, i.e.
the momentum (${\bf k}$), band ($\mu, \nu$),
spin ($ \sigma$) and Nambu degrees
of freedom. In the following, we consider only
interband transitions
$\mu \neq \nu$, and the limit of the dipole
approximation
${\bf q} \rightarrow {\bf 0}$ can be
performed without special care for
plasmonic excitations.
The evaluation of the trace of Eq.~\ref{nolispec}
is straightforward
with the knowledge of the unitary
transformation ${\cal U}_{{\bf k}\mu}$,
which diagonalizes $H$ and $\varrho(\epsilon)$.
In the following we discuss ${\cal U}_{{\bf k}\mu}$
for a bulk singlet
superconductor in the limit
of weak spin-orbit interaction $\lambda_{s.o.}$
and weak external field
because the phase sensitive contributions of
the optical response vanish if either
$\lambda_{s.o.}$ or the magnetic
field is zero.
For weak external magnetic field, the
eigenvalues are given by
\begin{equation}
E_{{\bf k}\mu}=\pm h \pm
{\cal E}_{{\bf k}\mu}\, ,
\label{eigens}
\end{equation}
where
\begin{equation}
{\cal E}_{{\bf k}\mu}=
\sqrt{\varepsilon_{{\bf k}\mu}^2 +\frac{1}{2}
{\rm tr} \left(\hat{\Delta}_{{\bf k}\mu}^\dagger
\hat{\Delta}_{{\bf k}\mu}
\right)}\, .
\end{equation}
$h$ is the absolute mangnitude of $\vec{h}$ and
${\rm tr}$ denotes the trace in
spin space. $\hat{\Delta}_{{\bf k}\mu}$ belongs
to one of the irreducible
representations of table~\ref{tab1}.
Analogously, the unitary transformation is
given by
\begin{equation}
{\cal U}_{{\bf k}\mu} ={\cal U}^\Delta_{{\bf k}\mu}\,
{\cal U}^h_{{\bf k}\mu}\, ,
\label{transcomb}
\end{equation}
where ${\cal U}^\Delta_{{\bf k}\mu}$ and
${\cal U}^h_{{\bf k}\mu}$
are the transformations which diagonalize
${\cal H}_{{\bf k}\mu}$
for $\vec{h}=0$ and $\vec{d}_{{\bf k}\mu}=0$,
respectively.
This is correct up to first order in
$\lambda_{s.o.}$ and $\vec{h}$.
The zero-field transformation
${\cal U}^\Delta_{{\bf k}\mu}$
can be expressed in terms of $(2\times2)$
matrices $\hat{u}_{{\bf k}\mu}$
and $\hat{v}_{{\bf k}\mu}$~\cite{SU91}:
\begin{equation}
{\cal U}^\Delta_{{\bf k}\mu}=
\left( \begin{array}{cc}
\hat{u}_{{\bf k}\mu} &\hat{v}_{{\bf k}\mu} \\
\hat{v}_{-{\bf k}\mu}^\ast &\hat{u}_{-{\bf k}\mu} ^\ast
\end{array} \right) \, ,
\end{equation}
where
\begin{equation}
\hat{u}_{{\bf k}\mu}=\left(
{\cal E}_{{\bf k}\mu}-\varepsilon_{{\bf k}\mu}
\right) /
{\cal E}_{{\bf k}\mu} \, \hat{\sigma}^o
\end{equation}
and
\begin{equation}
\hat{v}_{{\bf k}\mu}=
-\hat{\Delta}_{{\bf k}\mu}/{\cal E}_{{\bf k}\mu}\, .
\label{vmatrix}
\end{equation}
Similar to the gap function in
Eq.~\ref{gapexp}, these $(2 \times 2)$ matrices
are expanded in Pauli matrices, leading
to a singlet part $v^o_{{\bf k}\mu}$
and triplet part $\vec{v}_{{\bf k}\mu}$
of $\hat{v}_{{\bf k}\mu}$.
The transformation ${\cal U}^h_{{\bf k}\mu}$
is determined by a rotation
in spin space
\begin{equation}
{\cal U}^h_{{\bf k}\mu}= \left( \begin{array}{cc}
\exp\left(-i \vec{a}\cdot \vec{\sigma}\right) & 0 \\
0 & \exp\left(i \vec{a}\cdot \vec{\sigma}^\ast \right)
\end{array} \right) \, ,
\end{equation}
with rotation axis $ \vec{e}_a= -( \vec{e}_z
\times \vec{e}_h)/
| \vec{e}_z \times \vec{e}_h | $ and angle
$\cos(2a)=\vec{e}_z \cdot\vec{e}_h$
($\vec{a}=a \vec{e}_a$).
$\vec{e}_h$ is the unit vector in the
direction of $\vec{h}$.
Using this transformation, we can
diagonalize the Hamiltonian leading
to the quasiparticle spinor
\begin{equation}
\Psi_{{\bf k}\mu} = {\cal U}_{{\bf k}\mu}
\Phi_{{\bf k}\mu} \, .
\end{equation}
Here the components of $\Phi_{{\bf k}\mu}$
are the destruction operators of
the eigenstates with eigenvalues
given in Eq.(~\ref{eigens}).
In order to express the current operator
in terms of the Nambu spinors, we have
to consider the behavior of the matrix
element $\vec{j}_{{\bf k}\mu \nu}$
under the simultaneous transformation
${\bf k} \rightarrow -{\bf k} $ and
$(\mu,\nu) \rightarrow (\nu,\mu)$:
\begin{equation}
\vec{j}_{-{\bf k}\mu \nu} = p^{\mu \nu}_{{\bf k}}
\vec{j}_{{\bf k}\nu \mu} \, .
\end{equation}
Although the phase factors $p^{\mu \nu}_{{\bf k}}$
depend on the choice of the
phase of the Wannier functions, all observable
quantities like
$\chi_{\alpha \beta \gamma}({\bf q}, \omega) $
are independent of this choice.
Now, the current operator can be expressed
in terms of the Nambu spinors:
\begin{equation}
\vec{J}=\frac{1}{2} \sum_{{\bf k} \mu \nu}
\vec{j}_{{\bf k} \mu \nu}
\Psi^\dagger_{{\bf k} \mu}
\left( \begin{array}{cc}
\hat{\sigma}^o & 0 \\
0 & - p^{\mu \nu}_{{\bf k}} \hat{\sigma}^o
\end{array} \right)
\Psi_{{\bf k} \nu}\, ,
\label{currNam}
\end{equation}
and we find for the nonlinear spectral
function of Eq.(~\ref{nolispec}):
\begin{eqnarray}
\lefteqn{ I_{\alpha \beta \gamma}
(\epsilon,\epsilon ',\epsilon '')
=\frac{1}{8} \sum_{{\bf k} \mu \nu \kappa}
j^\alpha_{{\bf k} \mu \nu}
j^\beta_{{\bf k} \nu \kappa}
j^\gamma_{{\bf k} \kappa \mu} }
\nonumber \\
& & \times {\rm Tr}'
\left\{ {\cal M}_{{\bf k}}^{\mu \nu} \,
\varrho_{{\bf k}}^\nu(\epsilon) \,
{\cal M}_{{\bf k}}^{\nu \kappa} \,
\varrho_{{\bf k}}^\kappa(\epsilon ')\,
{\cal M}_{{\bf k}}^{\kappa \mu} \,
\varrho_{{\bf k}}^\mu(\epsilon '') \right\} \, .
\label{tr1}
\end{eqnarray}
Here, ${\cal M}_{{\bf k} }^{\mu \nu}$ and
$\varrho_{{\bf k}}^\nu(\epsilon)$ are
($4 \times 4$) matrices in Nambu and
spin space, whereby
\begin{equation}
{\cal M}_{{\bf k} }^{\mu \nu} =
{\cal U}_{{\bf k}\mu}^{\dagger}
\left( \begin{array}{cc}
\hat{\sigma}^o & 0 \\
0 & - p^{\mu \nu}_{{\bf k}} \hat{\sigma}^o
\end{array} \right)
{\cal U}_{{\bf k} \nu}
\label{MM}
\end{equation}
results from the transformation of
Eq.(~\ref{currNam}) into the quasiparticle
representation, where
$\varrho_{{\bf k}}^\nu(\epsilon)$ is diagonal.
Consequently, the trace ${\rm Tr}'$
has to be performed with respect to the
Nambu and spin degrees of freedom.
Nonlinear optics and in particular
nonlinear magneto-optics offers
a unique method to probe low-lying
excitations close to the Fermi-level
with optical photons, since SHG, which
involves three photons,
takes advantage of an additional degree
of freedom that is absent in linear
optics. Thus it allows to use conventional
monochromatic, intense, and
tunable pulse laser sources in the ps to
fs regime such as mode-locked
Ti-sapphire lasers. Besides, in nonlinear
interband optics, there is no
collective plasmon background which is
material insensitive and may
severely hamper the interpretation of
linear optical experiments.
Therefore, we restrict ourselves to a
special interband excitation process.
We consider the transition from the
initial state $i$ with energy
$E_i \approx -3 {\rm eV}$, below the
Fermi energy
to the intermediate state $s$ at the
Fermi level (which is the only
superconducting state) and to the
final state $f$
with energy $E_f \approx 3 {\rm eV}$
above $E_{\rm F}$, i.e.
we consider $\mu=f$, $\kappa=s$, and $\nu=i$.
A possible, but not necessary, origin
of the states $i$ and $f$ might be
due to the Mott-Hubbard splitting of
the hybridized Cu 3d$_{x ^2 - y ^2}$
and O 2p$_{x(y)}$ orbitals.
Since the intermediate state is the
only state with superconducting coherence,
we skip the band index of the matrices
$\hat{u}_{{\bf k}}$ and
$\hat{v}_{{\bf k}}$.
Performing finally the traces in
Eq.(~\ref{tr1}), we obtain:
\begin{equation}
\chi_{\alpha \beta \gamma} ( \omega)=
\chi^{(0)}_{\alpha \beta \gamma}( \omega)+
\chi^{(h)}_{\alpha \beta \gamma}( \omega)
+{\cal O}(h^2) \, ,
\end{equation}
where
\begin{equation}
\chi^{(0)}_{\alpha \beta \gamma}( \omega)
=\sum_{\bf k} \,j^\alpha_{{\bf k} f i}
\, j^\beta_{{\bf k} i s} \, j^\gamma_{{\bf k} s f}\,
\left( | u ^o_{\bf k}|^2 \, G_1({\bf k},\omega)
+ | v ^o_{\bf k}|^2 \, G_2({\bf k},\omega) \right)
\end{equation}
is the zero field susceptibility tensor
in second harmonic
generation within the superconducting state
which gives already
a contribution without spin-orbit interaction.
The nonvanishing tensor elements for
$\chi^{(0)}_{\alpha \beta \gamma}( \omega)$
are the same as in the normal state
($\alpha \beta \gamma \in
\{zzz, zxx,zyy,xzx,xxz,yzy,yyz\}$).
This results here from the transformational
properties of the three matrix
elements $ j^\alpha_{{\bf k} f i}
j^\beta_{{\bf k} i s} j^\gamma_{{\bf k} s f}$,
which transform as the
corresponding combination of
the coordinates $x_\alpha x_\beta x_\gamma$,
even if a single matrix element
e.g. $ j^\alpha_{{\bf k} f i}$ does not
transform like $x_\alpha$.
The functions $ G_{1(2)}({\bf k},\omega)$
result from the numerous
combinations of Fermi functions and energy
denominators which occur
by performing the traces in spin and Nambu
space.
$\chi^{(h)}_{\alpha \beta \gamma}( \omega)$
is the new contribution of the magneto-optical
Kerr effect in the
superconducting state.
This can be seen from the symmetry relations
of the magneto-optical
susceptibility
\begin{equation}
\chi^{(h)}_{\alpha \beta \gamma}( \omega)
=\sum_{\bf k} \,j^\alpha_{{\bf k} f i}
\, j^\beta_{{\bf k} i s} \, j^\gamma_{{\bf k}
s f}\,( v ^o _{\bf k} ) ^\ast
\, \vec{v}_{\bf k} \cdot \vec{e}_h \,
F({\bf k},\omega)\, ,
\end{equation}
which depends on the superconducting gap
function not only through
its magnitude but also through
$\Delta ^o_{\bf k}$ itselfe.
Due to the additional triplet part however,
the result is still gauge invariant.
The function $F({\bf k},\omega)$ correspond
to the $ G_{1(2)}({\bf k},\omega)$
for $\chi^{(h)}_{\alpha \beta \gamma}( \omega)$.
Considering a magnetic field parallel to the
x-axis (in plane), the nonvanishing
elements of $\chi ^{(h)}_{\alpha \beta \gamma}
( \omega)$ are:
$\alpha \beta \gamma \in
\{yyy,xxy,xyx,yxx,zzy,zyz,yzz\}$.
This results from the combination of the
transformation properties of the
normal state matrix elements and of the
symmetry sensitive term
$(v ^o _{\bf k}) ^\ast \vec{v}_{\bf k}
\cdot \vec{e}_h$.
Using Eq.(~\ref{vmatrix}),
the ${\bf k}$-dependence of the latter
results from
the irreducible representations given
in table~\ref{tab1}.
The above matrix elements lead to a rotation
of the polarization of the
incident light due to the interference of
the singlet and triplet
states at the surface of a superconductor.
Thus nonlinear magneto-optics,
unlike linear optical probes, provides
indeed an optical method to
discriminate different superconducting pairing
symmetries by exclusively
employing the effect of optical photons to
low energy excitations.
In the next paragraph, we discuss the
numerical results of the above model
bandstructure for the most realistic
experimental setups
and show how experiments can distinguish
between certain
symmetries of the gap function using
nonlinear magneto-optics.
\section{RESULTS}
In this section we discuss the numerical
results obtained for the tensor
elements $\chi^{(0)}_{zzz}$ of second
harmonic generation without
magnetic field and $\chi^{(h)}_{yzz}$
which gives rise to the
rotation of the polarization plane for
an applied magnetic field
parallel to the $x$-axis.
For simplicity we neglect the dispersion
of the initial and the final
state and consider solely the
${\bf k}$-dependence which results from
the superconducting gap function.
The momentum summations are performed within
the two dimensional
Brillouin zone using $81 \times 81$
${\bf k}$-points.
The calculations are performed for a magnetic
field of $9 \, {\rm T}$
(corresponding to a field induced band splitting
of 0.5 meV)
and a temperature of $1.5\, {\rm K}$.
The magnitude of the singlet part of the gap
function is assumed to be
$ 5 \, {\rm meV}$. Furthermore, the magnitude
of the dipole matrix elements
is estimated to be 10$^{-11}$ m.
All results presented here are not sensitive to
the specific set of parameters
chosen, but are typical for reasonable values of
the corresponding energy
scales. This was checked by systematically varying
the dependence of the
nonlinear susceptibility on the magnitude of the
gap function, the magnetic
field, the position of the initial and final states
$E_i$ and $E_f$, the
temperature and the linewidth broadening $\delta$.
In Figs.~\ref{fig1}(a) and (b), we show
$\omega ^2 {\rm Im}\chi^{(o)}_{zzz}(\omega)$ and
$\omega ^2 {\rm Im} \chi^{(h)}_{yzz}(\omega)$
for an isotropic $s$-wave.
For the conventional SHG, we find a line shape
similar to the real part of a
Lorentzian, which is typical for a three level
system discussed in this paper.
More interestingly, the fine structure of the
peak, shown in the inset of
Fig.~\ref{fig1}(a) clearly shows the energy
scale of the superconducting gap.
Comparing this behavior with the Kerr signal
$\omega ^2 {\rm Im} \chi^{(h)}_{yzz}(\omega)$
of Fig.~\ref{fig1}(b),
one finds that the interference of the singlet
and triplet pairing states
leads to a line shape with several pronounced
zeros and with a fine structure
that yields, besides the energy scale of the
superconducting gap, also
excitations which result from the magnetic
field splitting.
In all our calculations, this line shape was
exclusively observed for an
isotropic $s$-wave and can be considered as
a fingerprint of this symmetry.
In Figs.~\ref{fig2} and ~\ref{fig3}, the
corresponding results for
the anisotropic $s$-wave and the
$d_{x ^2 - y ^2}$-wave are shown.
Although the result for the conventional
SHG is similar to that
of the isotropic $s$-wave, a totally
different line shape of the Kerr
signal results.
This is due to the symmetry dependent
prefactor in $\chi^{(h)}_{yzz}(\omega)$
and can be used to discriminate these
two symmetries from the isotropic
$s$-wave.
Furthermore the fine-structure of these
two symmetries is very different
from the isotropic $s$-wave. Due to the
occurrence of nodes in the gap,
not only a peak but a whole broad band
between $3\, {\rm eV}$ and
$3.01\, {\rm eV}$ is observable. This range
is surprisingly given by twice the
superconducting gap magnitude.
Unfortunately, there are only slight
differences between the two
symmetries shown in Fig.~\ref{fig2}
and ~\ref{fig3}. This is due to the
similar ${\bf k}$-dependence of the triplet
part given in table~\ref{tab1}.
Only the fine structure of the peaks displayed
in the insets of
Fig.~\ref{fig2}(b) and ~\ref{fig3}(b) exhibits
a clear difference, where the
anisotropic $s$-wave has a clear zero at
$3.01\, {\rm eV}$ which is
more or less smeared out for the
$d_{x ^2 - y ^2}$ wave.
In this context it is of importance to
compare these results to the usual
nonlinear Kerr effect in the normal state
in an external applied field
which can be easily estimated using
previous results by
Pustogowa {\em et al.}~\cite{pustogowa}
as
\begin{equation}
\frac{\chi^{(h)}_{yzz,normal}}
{\chi^{(0)}_{zzz}}\;\approx\;
\frac{\lambda_{s.o}J(h)}{(\hbar\omega)^{2}}\;.
\end{equation}
Here, the incident photon energy $\hbar\omega$
at resonance is 3 eV, while
the energy splitting $J(h)$ caused by the
external magnetic field $h$ is
0.5 meV (see above). For this small value of $J(h)$,
the formula given in
appendix C of the paper by Pustogowa {\em et al.}
yields a linear dependence.
Thus, we find $\frac{\chi^{(h)}_{yzz,normal}}
{\chi^{(0)}_{yzz}}\;\approx\;
2.7\;\cdot\;10^{-6}$ due to the absence of a
spontaneous magnetization.
{}From this estimate, it follows that the
observability of the new
contribution to the nonlinear Kerr effect
is clearly guaranteed
for the anisotropic s-wave.
Although, the intensities of the other
symmetries are smaller
than the estimated value of the usual
nonlinear Kerr effect in both
the
normal and superconducting state, we believe
that this effect is
still
observable for the following reason:
Due to the neglect of the dispersion of the
states in our model
bandstructure,
the disapperance of the signal results from
cancellations of
contributions
of the order of magnitude of the anisotropic
s-wave in the ${\bf
k}$-summation.
This is due to the artificially high symmetry
of this bandstructure.
The simplified description of the high-T$_c$
materials is used
because the aim of this paper is to
demonstrate the strong
interdependence
of the lineshape of the nonlinear magneto-optical
susceptibility and
the
symmetry of the superconducting state.
A more realistic bandstructure immediately
leads to larger
intensities of
$\chi ^{(h)} _{yzz}$, while keeping the
characteristics of the line
shapes of the spectra.
In Fig.~\ref{fig4} we finally show our results
for the $d_{xy}$ symmetry.
Although, this symmetry does not seem to be
the most probable candidate
for the high-T$_c$ materials, it shows most
clearly the symmetry dependence
of the nonlinear magneto-optical Kerr effect.
In contrast to the
anisotropic $s$-wave and the $d_{x ^2 - y ^2}$
wave, there occurs a sign change
between the magnetic and nonmagnetic optical
spectrum.
This is observable since the sign of the Kerr
spectrum determines the direction
of the rotation of the polarization axis,
i.e. the Kerr angle.
Furthermore, the satellites shown in the
inset of Fig.~\ref{fig4}(a) and (b)
cover only the range from $3\, {\rm eV}$
to $3.005 \, {\rm eV}$, i.e.
only one times the gap magnitude.
For the existence of a finite Kerr signal,
it is necessary to break
time reversal symmetry and to apply an external
magnetic field.
This enables one to keep any direction of the
field fixed and to study
the anisotropy of the effects discussed in
this paper.
However, due to the strong but short-ranged
antiferromagnetic correlations
it might also be possible to take advantage
of the locally broken time-reversal
symmetry of the high-T$_c$ materials.
Since a finite Kerr signal is expected for
certain long-range ordered
antiferromagnets~\cite{fr"ohlich}, a
pump-and-probe experiment
(on a time scale faster than the average
lifetime
of the local spin configurations
$\tau_{\rm spin} \approx 10 ^{-12} -
10 ^{-13} \, {\rm s}$)
could be able to resolve the influence
of the neighboring spins on the
site which is excited by the optical
excitation.
Furthermore, for a practical realization
of the experiment, one has to take into
account the possible heating effects of the
incoming light on the sample, which
might lead to a local disappearance of the
superconducting state.
In view of the comparable excitation energies
of the nonlinear magneto-optical
Kerr effect in ferromagnets, we believe that
this effect is of minor importance,
since for fs laser pulses sample heating is
of the order of 10 K and thus
negligible. This simple estimate is readily
obtained from the comparison of
laser heating with ns pulses~\cite{vaterlaus}
yielding intensities of
100 MWm$^{-2}$ and temperature rises of the
order of 100 K and typical fs
measurements of the nonlinear
Kerr-effect~\cite{theo} which operate at laser
powers as low as 100 mW focused on spot
diameters of 100 $\mu $m.
In conclusion, we presented a theory for
the nonlinear magneto-optical response
of superconductors. The surface sensitivity
of this experiment is of
particular interest for the simultaneous
occurrence of singlet and triplet pairing
amplitudes.
Therefore, a suitable material for this
experiment is the
Bi$_2$Sr$_2$CaCu$_2$O$_8$ compound, where
almost no reconstruction
of the cleaved surface (Bi-O layer) occurs,
and where the existence of
the superconducting state in the upper
CuO$_2$ layer
was clearly shown in photoemission
experiments.~\cite{SDW93}.
Since, due to difficulties in the manufacture
of the tunneling contacts,
all corner junction experiments where so far
performed with
YBa$_2$Cu$_3$O$_{7-\delta}$ samples, this
gives also insight into the
{\em symmetry} of another class of cuprate
compounds.
Note that furthermore the surface sensitivity
does not depend on the
actual depth beneath the
surface where superconductivity starts,
since the electronic symmetry
of the superconducting state surface is
of relevance in this context
which may or may not be perfectly identical
to the physical surface.
On the other hand the new nonlinear
magneto-optical effect proposed in
this paper may also be of considerable
importance for interfaces between
$s$- and $d$-wave superconductors.
Furthermore, we performed a group-theoretical
classification of the irreducible
representations of the gap function at the
surface of a tetragonal system.
Based on this theory, we showed that the
interference of singlet and triplet
pairing states in a magnetic field leads to
a new contribution to the
nonlinear magneto-optical Kerr signal which
is sensitive to certain
{\em symmetries} of the superconducting order
parameter rather than only to its
{\em magnitude}.
This enables us to give the basic line shapes
of the corresponding optical
spectra and to show that it is possible to
discriminate an isotropic $s$-wave
and a $d_{xy}$ wave from the anisotropic
$s_{x ^2 + y ^2}$ and
$d_{x ^2 - y ^2}$ waves. Unfortunately, it
is not possible to discriminate
between the two latter symmetries which seem
to be the most probable
symmetries of the high-$T_c$ materials.
Nevertheless, we believe that the nonlinear
optic can yield information
complementary to the tunneling experiments
and might be also of importance
in view of the application on heavy fermion
superconductors, where it was
also manifested that the superconducting state
is anomalous~\cite{B84,B90,K92}.
\acknowledgments
We would like to thank Prof. K.-H. Bennemann
for stimulating discussions and his continued
interest in this work.
|
\section{Introduction}
Affine Toda field theories are two-dimensional models described by the
Euclidean action
\begin{equation}
S = \int d^2x\left(\frac{1}{2}\partial_\mu\phi\cdot\partial_\mu\phi
+\frac{m^2}{\beta^2}\sum_{a=0}^r n_a e^{\beta \alpha^{(a)}\cdot\phi}\right).
\end{equation}
The $r$-dimensional vectors $\{\alpha^{(a)}\}$ are the simple roots of an
affine Kac-Moody algebra~\cite{Kac} and $\{n_a\}$ are positive integers
depending on the algebra and satisfying
\begin{equation}
\sum_a n_a \alpha_a = 0, \qquad n_0 = 1.
\end{equation}
The field $\phi$ is a set of $r$ real scalar components. Finally, $m$ and
$\beta$ are a mass scale parameter and the coupling constant.
The Coxeter number is the positive integer $h=n_0+\cdots +n_r$.
Under the transformation $T:\alpha\to 2\alpha/|\alpha|^2$, the lattice of the
simple roots transforms into the lattice of another affine algebra.
The invariant algebras are called self-dual; they belong to
the untwisted a-d-e series $a_n^{(1)}$, $d_n^{(1)}$, $e_n^{(1)}$
and to the twisted series $a_{2n}^{(2)}$. The other algebras are the pairs
$(b_n^{(1)}, a_{2n-1}^{(2)})$, $(c_n^{(1)}, d_{2n-1}^{(2)})$,
$(g_2^{(1)}, d_4^{(3)})$, and $(f_4^{(1)}, e_6^{(2)})$; they are invariant
under $T$.
At the classical level, affine Toda theories have no coupling; $\beta$ can be
scaled away, the spectrum is proportional to $m$, independent of $\beta$ and
moreover it is given by simple universal formulae in terms of the Coxeter
number.
The interest of the classical theory is that the field equations of motion
admit a Lax pair and therefore
there is an infinite hierachy of conserved currents with increasing spin.
At the quantum level, this property is inherited in the form of a factorized
S-matrix. The dependence on $\beta$ which plays the role of Planck's constant
becomes non trivial; on the other hand the parameter $m$ becomes unphysical
due to renormalization effects and only mass ratios are observables.
Since the S-matrix is expected to be factorizable, its explicit form may be
sought. One can make a guess and impose physical constraints like unitarity
or crossing symmetry
and the additional bootstrap principle. In the case of the self-dual
theories, perturbation theory suggests that the mass ratios do not
renormalize. Indeed, the bootstrap equations close on an ansatz for the
S-matrix based on the tree
level spectrum and on the fusings allowed by the three-point
couplings~\cite{Braden,Mussardo1}.
Perturbation theory and the structure of the bootstrap suggest conjectured
expressions
for the exact $\beta$ dependence of the S-matrix which show a remarkable
duality
between weak and strong coupling in terms of the transformation $\beta\to
4\pi/\beta$.
As pointed out in~\cite{Mussardo2}, for the non self dual pairs the picture is
more complicated. Mass ratios deform already at the lowest order of perturbation theory
and the simplest ansatz for the S-matrix fails.
However, a non trivial solution to the bootstrap equations can be found with
the feature of predicting $\beta$ dependent mass
ratios~\cite{Delius,CorriganA}.
The predictions are then formally the same as in the classical
theory, but in terms of a ``renormalized'' Coxeter number $H(\beta)$.
Again, the explicit non perturbative form of $H(\beta)$ is not known.
The simplest conjecture~\cite{Dorey}, consistent with low order perturbation
theory~\cite{Cho} and current algebra~\cite{Kausch}
predicts a new kind of duality. Under $\beta\to 4\pi/\beta$ the S-matrices of
the pair members get exchanged. Hence, the strong coupling regime in one theory
should be given by the weak coupling regime in the other.
In~\cite{Watts} a Monte Carlo study of duality in the pair
$(g_2^{(1)}, d_4^{(3)})$ was performed
by mean of the Metropolis algorithm. The authors determined the mass ratio in
the $g_2^{(1)}$ theory over a wide range of couplings and they did find
agreement
with the duality conjecture. Specifically, they checked that the mass ratio
in $g_2^{(1)}$ ranged
between its classical values and the classical value of $d_4^{(3)}$ .
In this paper I carried over the above simulation on larger lattices with
higher statistics in order to pin down the precise dependence on $\beta$.
Moreover, I have used the Hybrid Monte Carlo algorithm~\cite{Duane}. Finally,
I have extended the simulation to the $d_4^{(3)}$ theory in order to have a
complete picture.
The plan of the paper is the following: Section~\ref{pair} describes the pair
$(g_2^{(1)}, d_4^{(3)})$ ; Section~\ref{oneloop} shows the one loop
deformations of the mass
ratios;
Section~\ref{simulation} gives some detail on the numerical simulation;
finally,
in Section~\ref{results} the results are discussed.
\section{The dual pair ($g_2^{(1)}$, $d_4^{(3)}$)}
\label{pair}
The pair ($g_2^{(1)}$, $d_4^{(3)}$) has $r=2$ and its action is
\begin{equation}
S = \int d^2x \left\{
\frac{1}{2} \partial_\mu\phi\cdot \partial_\mu\phi
+\frac{m^2}{\beta^2}\sum_{a=0}^2
n_a \exp(\beta\ \alpha_a\cdot\phi)
\right\},\qquad \phi = (\phi_1, \phi_2).
\end{equation}
The integers $n_a$ and the simple roots are
\begin{eqnarray}
g_2^{(1)} &:& n = \{2,3,1\},\qquad
\alpha = \left\{\left(\sqrt{2},0\right),
\left(-\frac{1}{\sqrt{2}},\frac{1}{\sqrt{6}}\right),
\left(-\frac{1}{\sqrt{2}},-\sqrt{\frac{3}{2}}\right)\right\}; \\
d_4^{(3)} &:& n = \{2,1,1\},\qquad
\alpha = \left\{\left(\sqrt{2},0\right),
\left(-\frac{3}{\sqrt{2}},\sqrt{\frac{3}{2}}\right),
\left(-\frac{1}{\sqrt{2}},-\sqrt{\frac{3}{2}}\right)\right\}.
\end{eqnarray}
The two sets of roots are related by the duality
$\alpha\to 2\alpha/|\alpha|^2$.
The two corresponding models are very different at the tree level.
The explicit expansion of the mass-potential exponential
term up to the fourth order for the $g_2^{(1)}$ model is
\begin{eqnarray}
\lefteqn{V(\phi_1,\phi_2) = m^2\left(3\,{\phi_1^2} + {\phi_2^2}\right) +} &&
\nonumber \\
&& + m^2\beta {{9\,{\phi_1^3} - 9\,\phi_1\,{\phi_2^2} - 2\,{\sqrt{3}}\,
{\phi_2^3}}\over
{9\,{\sqrt{2}}}} + \\
&& + m^2\beta^2 {{27\,{\phi_1^4} + 18\,{\phi_1^2}\,{\phi_2^2} +
8\,{\sqrt{3}}\,\phi_1\,{\phi_2^3} + 7\,{\phi_2^4}}\over {72}} +
O(\beta^3). \nonumber
\end{eqnarray}
For the $d_4^{(3)}$ model we utilize the tree level mass eigenstates by
transforming
the fields
\begin{equation}
\phi\to R\phi ,\qquad R = \left(\begin{array}{cc}
\cos\theta & \sin\theta \\
-\sin\theta & \cos\theta
\end{array}\right), \qquad \theta = \frac{5 \pi}{12}
\end{equation}
and obtain the expansion
\begin{eqnarray}
\lefteqn{V(\phi_1,\phi_2) = m^2\left( (3 - {\sqrt{3}})\,{\phi_1^2} +
(3+{\sqrt{3}})\,{\phi_2^2} \right) + } && \nonumber\\
&& + m^2\beta\,\left( {{{\phi_1^3}}\over 2} - {{{\sqrt{3}}\,{\phi_1^3}}
\over 2} +
{{3\,{\phi_1^2}\,\phi_2}\over 2} - {{{\sqrt{3}}\,{\phi_1^2}\,\phi_2}
\over 2} +
{{3\,\phi_1\,{\phi_2^2}}\over 2} + {{{\sqrt{3}}\,\phi_1\,{\phi_2^2}}
\over 2} +
{{{\phi_2^3}}\over 2} + {{{\sqrt{3}}\,{\phi_2^3}}\over 2} \right) + \\
&& +m^2\beta^2 \,\left( {{7\,{\phi_1^4}}\over 8} -
{{5\,{\phi_1^4}}\over {4\,{\sqrt{3}}}} + {{{\phi_1^3}\,\phi_2}\over 2} -
{{{\phi_1^3}\,\phi_2}\over {{\sqrt{3}}}} + {{3\,{\phi_1^2}\,{\phi_2^2}}
\over 4} +
{{\phi_1\,{\phi_2^3}}\over 2} + {{\phi_1\,{\phi_2^3}}\over {{\sqrt{3}}}}
+
{{7\,{\phi_2^4}}\over 8} + {{5\,{\phi_2^4}}\over {4\,{\sqrt{3}}}}
\right) + O(\beta^3) \nonumber
\end{eqnarray}
As one can see, the sets of possible fusings are completely different and
duality is far from being obvious.
The classical mass ratios are
\begin{equation}
\left.\frac{m_2}{m_1}\right|_{g_2^{(1)}} = \sqrt{3},\qquad
\left.\frac{m_2}{m_1}\right|_{d_4^{(3)}} = \sqrt{\frac{\sqrt{3}+1}
{\sqrt{3}-1}} =
\frac{\sqrt{3}+1}{\sqrt{2}}
\end{equation}
which agree with the general formula in terms of the Coxeter number $h$
(6 for $g_2^{(1)}$ , 12 for $d_4^{(3)}$)
\begin{equation}
\frac{m_2}{m_1} = \frac{\sin(2\pi/h)}{\sin(\pi/h)} = 2 \cos(\pi/h)
\end{equation}
The duality conjecture states that the correct quantum ratio
$g_2^{(1)}$ is given by substituting $h\to H(\beta)$ in the model $g_2^{(1)}$
and
$h\to H(4\pi/\beta)$ in the $d_4^{(3)}$ model.
The form of $H(\beta)$ is constrained but not fixed by perturbation theory
and the conjectured expression is
\begin{equation}
H(\beta) = 6+\frac{\beta^2/2\pi}{1+\beta^2/12\pi}.
\end{equation}
Let us clarify these statements by considering the one loop mass ratios.
\section{One loop mass ratios}
\label{oneloop}
Let us denote the three diagrams of Figures (I-II-III) by
\begin{equation}
\Gamma^{(1)}_{abc}, \quad \Gamma^{(2)}_{abcd}, \quad \Gamma^{(3)}_{abcd}
\end{equation}
where $a$, $b$, $c$ and $d$ are particle labels in the range $\{1,2\}$.
The mass ratio is observable since renormalization amounts to a normal
ordering of the
exponentials and its effect is a redefinition of the bare mass.
We must check that in a bare renormalization scheme all the divergent tadpole
graphs cancel.
Let us utilize dimensional regularization and let us introduce
\begin{equation}
Z_i = \int \frac{d^dp}{(2\pi)^d}\frac{1}{p^2+m_i^2} .
\end{equation}
The pole part of $Z_i$ is mass independent, hence the cancellation is a
matter of couplings. At the
one loop level the mixed propagator corrections are irrelevant and we can
restrict to the diagonal
ones. Let us write the interaction lagrangian in the form
\begin{equation}
V(\phi_1,\phi_2) = \frac{1}{2}(m_1^2\phi_1^2 + m_2^2\phi_2^2) +
V_{111}\phi_1^3 +
V_{112}\phi_1^2\phi_2 + \cdots .
\end{equation}
Then, the corrections to the propagator of particle 1 are
\begin{eqnarray}
\Gamma^{(1)}_{111} &=& -12\ V_{1111}\ Z_1,\nonumber\\
\Gamma^{(1)}_{112} &=& -2\ V_{1122}\ Z_2,\nonumber\\
\Gamma^{(2)}_{1111} &=& 18\ V_{111}^2\ Z_1\ m_1^{-2}, \\
\Gamma^{(2)}_{1112} &=& 6\ V_{111}\ V_{122}\ Z_2\ m_1^{-2},\nonumber\\
\Gamma^{(2)}_{1121} &=& 2\ V_{112}^2\ Z_1\ m_2^{-2},\nonumber\\
\Gamma^{(2)}_{1122} &=& 6\ V_{112}\ V_{222}\ Z_2\ m_2^{-2}. \nonumber
\end{eqnarray}
The corrections to the propagator of particle 2 are obtained by exchanging
the 1 and 2 labels.
If we denote the full divergent correction by
\begin{equation}
\delta m_1^2 = \Gamma^{(1)}_{111}+\Gamma^{(1)}_{112}+\Gamma^{(2)}_{1111}+
\Gamma^{(2)}_{1112}+\Gamma^{(2)}_{1121}+\Gamma^{(2)}_{1122}
\end{equation}
then the desired cancellation is equivalent to the condition
\begin{equation}
\frac{\delta m_1^2}{m_1^2}=\frac{\delta m_2^2}{m_2^2}
\end{equation}
which is indeed satisfied by the couplings of the two theories which can be
read in the
expansions of the previous section.
Besides the consistency check, let us turn to the mass ratio deformation.
At one loop, we must determine the quantity
\begin{equation}
\delta\ \frac{m_1^2}{m_2^2} = \frac{m_1^2}{m_2^2} \left(
\frac{\delta m_1^2}{m_1^2}-\frac{\delta m_2^2}{m_2^2}
\right).
\end{equation}
Let us introduce the finite integral
\begin{equation}
Z_{ij}(p^2) = \int \frac{d^2 q}{(2\pi)^2}
\frac{1}{(q^2+m_i^2)((q+p)^2+m_j^2)}.
\end{equation}
Then the finite contributions to the propagators of particle 1 are
\begin{eqnarray}
\Gamma^{(3)}_{1111} &=& 18 V_{111}^2 Z_{11}(p^2),\nonumber\\
\Gamma^{(3)}_{1112} &=& 4 V_{112}^2 Z_{12}(p^2),\\
\Gamma^{(3)}_{1122} &=& 2 V_{122}^2 Z_{22}(p^2).\nonumber
\end{eqnarray}
Evaluation on the tree mass shell gives
\begin{equation}
-\delta m_1^2 = 18 V_{111}^2 Z_{11}(-m_1^2) + 4 V_{112}^2 Z_{12}(-m_1^2)
+2 V_{122}^2 Z_{22}(-m_1^2)
\end{equation}
with analogous expressions for the particle 2.
We need only the following particular values
\begin{eqnarray}
Z_{ii}(-m_i^2) &=& \frac{1}{4\sqrt{3}}\frac{1}{m_i^2}, \\
Z_{ij}(-m_i^2) &=& \frac{1}{2\pi\sqrt{m_j^2(m_j^2-4m_i^2)}}\mbox{ArcTanh}
\sqrt{\frac{m_j^2-4m_i^2}{m_j^2}}
\end{eqnarray}
and the final result is
\begin{equation}
g_2^{(1)}: \qquad \delta\ \frac{m_1^2}{m_2^2} = \frac{1}{12\sqrt{3}} \beta^2
+ O(\beta^4); \qquad
d_4^{(3)}: \qquad \delta\ \frac{m_1^2}{m_2^2} = -\frac{1}{16} \beta^2 +
O(\beta^4).
\end{equation}
The renormalized Coxeter number is thus
\begin{equation}
g_2^{(1)}: \qquad H(\beta) = 6 + \frac{\beta^2}{2\pi} + O(\beta^4); \qquad
d_4^{(3)}: \qquad H(\beta) = 12 - \frac{9\beta^2}{2\pi} + O(\beta^4)
\end{equation}
and a consistent, simple and natural conjecture is
\begin{equation}
g_2^{(1)}: \qquad H(\beta) = H_0(\beta); \qquad
d_4^{(3)}: \qquad H(\beta) = H_0(4\pi/\beta)
\end{equation}
where
\begin{equation}
H_0(\beta) = 6 + \frac{\beta^2/2\pi}{1+\beta^2/12\pi} \qquad 6<H_0<12.
\end{equation}
The result for $g_2^{(1)}$ is in agreement with that quoted by~\cite{Delius}.
The result for $d_4^{(3)}$ gives perturbative support to the duality
conjecture
$\beta\to4\pi/\beta$.
We remark that the discrepancy with~\cite{Cho} is due to the fact that they
use the
form of $H_0$ which is correct for the simply laced models.
\section{Details of the simulation}
\label{simulation}
The lattice action for the pair ($g_2^{(1)}$, $d_4^{(3)}$) expressed in terms
of pure numbers is
\begin{equation}
S_{\rm Toda} = \sum_{n\in\rm sites}\left\{
\frac{1}{2}\sum_{\mu=1,2}
(\phi_{n+\mu}-\phi_n)^2 + \frac{m^2}{\beta^2}\sum_{a=1}^3
n_a \exp(\beta\ \alpha_a\cdot\phi)
\right\},\qquad \phi = (\phi_1, \phi_2).
\end{equation}
I have simulated the Toda theory with the Hybrid Monte Carlo algorithm
(see~\cite{Duane} for
the details).
Let us consider the extended action
\begin{eqnarray}
S &=& S_p + S_{\rm Toda}, \\
S_p &=& \frac{1}{2} \sum_n \pi_n\cdot \pi_n, \qquad \pi = (\pi_1, \pi_2).
\end{eqnarray}
The free parameter of the algorithm are $N_{hmc}$ and $\epsilon$. The first
is the number of molecular dynamics steps. The second is the time step in the
integration of the equations of motion
\begin{eqnarray}
\dot{\phi}_n &=& \pi_n, \\
\dot{\pi}_n &=& \sum_\mu (\phi_{n+\mu}-2\phi_n + \phi_{n-\mu}) -
\frac{m^2}{\beta} \sum_a n_a \alpha_a \exp(\beta\alpha_a\cdot\phi_n).
\end{eqnarray}
The vacuum expectation value of the field is a non physical quantity.
However, it is interesting to measure it since it is an indicator of
thermalization and also because it is in a sense a dynamic minimum of the
Toda potential.
Mass ratios can be determined by studying the eigenvalues of the two point
function
\begin{equation}
\langle 0 | \Phi_i(0)\Phi_j(\tau)| 0 \rangle - \langle 0 | \Phi_i| 0 \rangle
\langle 0 | \Phi_j| 0 \rangle
\end{equation}
where $t$ is the lattice time ranging from $0$ to $T$ and the wall field
$\Phi_i(t)$ is obtained by averaging $\phi$ over space.
\section{Results}
\label{results}
I have used a $80^2$ lattice for all $\beta$s because the correlation lenght
may be adjusted
by varying $m$. The continuum mass ratio is independent
of the bare mass $m$. However, on a finite lattice, it must be chosen
in order to have correlation lenghts large with respect to one lattice
spacing and small compared to the lattice size. This is the correct
procedure which minimizes discretization and finite size corrections.
Thanks to the work of~\cite{Watts} I had good values in the case of the
$g_2^{(1)}$ theory. In the other model, I started with the same values of $m$
adjusting them for some $\beta$.
I utilized different measurement of the wall-wall two-point
function for each bin in the separation $\tau$. This is necessary in
order to avoid strong correlation between data.
Table~\ref{table1} shows the Hybrid Monte Carlo parameters which we
found to be optimal for each couple $(\beta,m)$. The time step $\epsilon$
must be reduced almost exponentially as $\beta$ is increased. This is
reasonable since at larger $\beta$ the potential profile becomes steeper.
Table~\ref{table2} shows the measure of $\langle 0 |
\phi_i | 0 \rangle$ which can be useful as a check
of the code and which is needed in order to subtract the two-point function.
Table~\ref{table3} shows the two lattice masses, their ratio and the
conjectured prediction.
Finally, tables~\ref{table4}, \ref{table5}, \ref{table6} show the same results
in the case of the $d_4^{(3)}$ model.
Figures I-II-III show the self energy diagrams which are needed in order to
compute the one loop mass ratio deformations.
Figure IV shows a summary plot of the measured mass ratios in the two
models together with the conjectured ones and the asymptotic values holding
in the classical limit.
\section{Conclusions}
In this paper I have investigated numerically the conjectured duality
in the pair $(g_2^{(1)}, d_4^{(3)})$ of non simply laced affine Toda theories.
I have shown
that the $\beta$ dependence of the mass ratios in $g_2^{(1)}$ does follows the
behaviour conjectured in~\cite{Watts} and that the data of $d_4^{(3)}$
agree with the $\beta\to4\pi/\beta$ duality.
As in the case of more realistic field theories like QCD, the numerical
approach could be useful in studying other interesting features of quantum
Toda theories.
For instance, one could try to find direct evidence of the boundary bound
states which appears when the theory is restricted to a
half-line~\cite{CorriganB}; work is in progress on this topic.
Moreover, it could be valuable a non perturbative study of the solitons
which appear at imaginary $\beta$, and which suggests that a unitary theory
can ultimately be found by restricting the state space of the
hamiltonian~\cite{Hollowood};
their stability is indeed still questionable~\cite{Sasaki}.
\section{Acknoledgements}
I gratefully acknowledge G.~M.~T.~Watts and R.~A.~Weston for
useful suggestions and interest.
|
\section{Introduction}
One of the first results in the mathematics of computation, which underlies
the subsequent development of much of theoretical computer science,
was the distinction between computable and non-computable functions
shown in papers of Church [1936], Turing [1936], and Post [1936].
Central to this result is Church's thesis,
which says that all computing devices can be simulated by a
Turing machine. This thesis greatly simplifies the study of
computation, since it reduces the potential field of study from
any of an infinite number of potential computing devices to
Turing machines. Church's thesis is not a
mathematical theorem; to make it one would require a
precise mathematical description of a computing device.
Such a description, however, would leave open the possibility of
some practical computing
device which did not satisfy this precise mathematical description, and thus
would make the resulting
mathematical theorem weaker than Church's original thesis.
With the development of practical computers, it has become apparent
that the distinction between computable and non-computable functions
is much too coarse; computer scientists are now interested in
the exact efficiency with which specific functions can be computed. This
exact efficiency, on the other hand, is too precise a quantity to
work with easily. The generally accepted
compromise between coarseness and precision distinguishes
efficiently and inefficiently computable functions by whether
the length of the computation scales polynomially or superpolynomially
with the input size. The class of problems which can be
solved by algorithms having a number of steps polynomial in the
input size is known as~P.
For this classification to make sense, we need it to be machine-independent.
That is, we need to know that whether a function
is computable in polynomial time is independent of the kind of computing
device used. This corresponds to the following quantitative
version of Church's thesis, which
Vergis et al.\ [1986] have called the ``Strong Church's Thesis'' and
which makes up half of the ``Invariance Thesis'' of van Emde Boas [1990].
{\sc Thesis {\rm (Quantitative Church's thesis).}}
{\it
Any physical computing
device can be simulated by a Turing machine in a number of steps
polynomial in the resources used by the computing device.}
In statements of this thesis, the Turing machine is sometimes augmented with
a random number generator, as it has not yet been determined whether there
are pseudorandom number generators which can efficiently simulate truly
random number generators for all purposes.
Readers who are not comfortable with Turing machines may think instead
of digital computers having an amount of memory that grows linearly with
the length of the computation, as these two classes of computing machines
can efficiently simulate each other.
There are two escape clauses in the above thesis. One of these is
the word ``physical.'' Researchers
have produced machine models that violate the
above quantitative Church's thesis,
but most of these have been ruled out by some reason for why they are
not ``physical,'' that is, why they could not be built and made
to work. The other escape clause in the above thesis
is the word ``resources,'' the meaning of which is not completely specified
above. There are generally two resources which limit the ability of digital
computers to solve large problems: time (computation steps)
and space (memory). There are more resources pertinent to analog
computation; some proposed analog machines that seem able
to solve NP-complete problems in polynomial time have required the
machining of exponentially precise parts, or an exponential amount of energy.
(See Vergis et al.\ [1986] and Steiglitz [1988]; this issue is also
implicit in the papers of Canny and Reif [1987] and Choi et al.\ [1995]
on three-dimensional shortest paths.)
For quantum computation, in addition to space and time,
there is also a third potentially important resource, precision. For a
quantum computer to work, at least in any currently envisioned
implementation, it must be able to make changes in the quantum states
of objects (e.g., atoms, photons, or nuclear spins). These
changes can clearly not be perfectly accurate, but must contain some
small amount of
inherent imprecision. If this imprecision is constant (i.e., it does
not depend on the size of the input), then it is not
known how to compute any functions in polynomial time on a
quantum computer that cannot also be
computed in polynomial time on a classical computer with a random
number generator. However,
if we let the precision grow polynomially in the input size (that is,
we let the number of {\em bits} of precision grow logarithmically in
the input size), we appear
to obtain a more powerful type of computer. Allowing the same polynomial
growth in precision does not appear to confer extra computing power to
classical mechanics, although allowing exponential growth in precision does
\cite{HaSi,VeStDi}.
As far as we know, what
precision is possible in quantum state manipulation is dictated not by
fundamental physical laws but by the properties of the materials and
the architecture with which a quantum computer is built. It is
currently not clear which architectures, if any, will give high precision,
and what this precision will be. If the precision
of a quantum computer
is large enough to make it more powerful than a classical computer,
then in order to understand its potential
it is important to think of precision as a resource that can vary.
Treating the precision as a large constant (even though it is almost
certain to be constant for any given machine) would be comparable to
treating a classical digital computer as a finite automaton --- since
any given computer has a fixed amount of memory, this view
is technically correct; however, it is not particularly useful.
Because of the remarkable effectiveness of our mathematical models of
computation, computer scientists have tended to forget that computation
is dependent on the laws of physics. This can be seen in the
statement of the quantitative Church's thesis in van Emde Boas [1990], where
the word ``physical'' in the above phrasing is replaced with the word
``reasonable.'' It is difficult to imagine any definition of
``reasonable'' in this context which does not mean ``physically realizable,''
i.e., that this computing machine could actually be built and would work.
Computer scientists have become convinced of the truth of the
quantitative Church's thesis through the failure of all proposed
counter-examples. Most of these proposed counter-examples have
been based on the laws of classical mechanics; however, the
universe is in reality
quantum mechanical. Quantum mechanical objects often behave
quite differently from how our intuition, based on classical mechanics, tells
us they should. It thus seems plausible that the natural
computing power of classical mechanics corresponds to Turing
machines,\footnote{I believe that this question has not yet been settled
and is worthy of further investigation. See
Vergis et al.\ [1986], Steiglitz [1988], and Rubel [1989]. In particular,
turbulence seems a good
candidate for a counterexample to the quantitative Church's thesis
because the non-trivial dynamics on many length scales may make it
difficult to simulate on a classical computer.}
while the natural computing power of
quantum mechanics might be greater.
The first person to look at the interaction between computation and
quantum mechanics appears to have been Benioff [1980, 1982a, 1982b].
Although he did not ask whether quantum mechanics conferred extra power to
computation, he showed that reversible unitary evolution was sufficient
to realize the computational power of a Turing machine,
thus showing that quantum mechanics is at least as powerful computationally
as a classical computer. This work was fundamental in making
later investigation of quantum computers possible.
Feynman [1982,1986] seems to have been the first to suggest that quantum
mechanics might be more powerful computationally than a Turing machine.
He gave arguments as to why quantum mechanics might be
intrinsically expensive computationally to simulate on a classical
computer. He also raised the possibility of using a
computer based on quantum mechanical principles to avoid this problem, thus
implicitly asking the converse question: by using quantum mechanics in a
computer can you compute more efficiently than on a classical computer?
Deutsch [1985, 1989] was the first to ask this question explicitly.
In order to study this question,
he defined both quantum Turing machines and quantum circuits and
investigated some of their properties.
The question of whether using quantum mechanics in a computer allows one to
obtain more computational power was more recently addressed by
Deutsch and Jozsa [1992] and Berthiaume and Brassard [1992a, 1992b].
These papers showed that there are problems which quantum computers
can quickly solve exactly, but that classical computers can only solve
quickly with high probability and the aid of a random number generator.
However, these papers did not show how to solve any problem
in quantum polynomial time that was not already known to be solvable in
polynomial time with the aid of a random number generator,
allowing a small probability of error;
this is the characterization of the complexity class BPP, which
is widely viewed as the class of efficiently solvable problems.
Further work on this problem was stimulated by Bernstein and Vazirani [1993].
One of the results contained in
their paper was an oracle problem (that is, a problem involving a
``black box'' subroutine that the computer is allowed to perform,
but for which no code is accessible) which can be done in polynomial
time on a quantum Turing machine but which requires super-polynomial time
on a classical computer. This result was improved
by Simon [1994], who gave a much simpler construction of an
oracle problem which takes polynomial time on a quantum computer but
requires {\em exponential} time on a classical computer. Indeed,
while Bernstein and Vaziarni's problem appears contrived, Simon's
problem looks quite natural.
Simon's algorithm inspired the work presented in this paper.
Two number theory problems which have been studied extensively but
for which no polynomial-time algorithms have yet been discovered are
finding discrete logarithms and factoring integers
[Pomerance 1987, Gordon 1993, Lenstra and Lenstra 1993, Adleman and
McCurley 1995]. These problems
are so widely believed to be hard that several cryptosystems based on their
difficulty have been proposed, including the widely used
RSA public key cryptosystem developed by Rivest, Shamir, and Adleman [1978].
We show that these
problems can be solved in polynomial time on a quantum computer
with a small probability of error.
Currently, nobody knows how to build a quantum computer, although it seems as
though it might be possible within the laws of quantum mechanics.
Some suggestions have been made as to possible designs for such computers
[Teich et al.\ 1988, Lloyd 1993, 1994, Cirac and Zoller 1995,
DiVincenzo 1995, Sleator and Weinfurter 1995, Barenco et al.\ 1995b,
Chuang and Yamomoto 1995], but there will
be substantial difficulty in building any of these
\cite{Land1,Land2,Unru,ChLaShZu,PaSuEk}. The most difficult
obstacles appear to involve the decoherence of quantum superpositions
through the interaction of the computer with the environment, and
the implementation of quantum state transformations with enough precision
to give accurate results after many computation steps.
Both of these obstacles become more difficult as the size of the
computer grows, so it may turn out to be possible to
build small quantum computers, while scaling up to machines large enough
to do interesting computations may present fundamental difficulties.
Even if no useful quantum computer is ever built, this research does
illuminate the problem of simulating quantum mechanics on a classical
computer. Any method of doing this for an arbitrary Hamiltonian would
necessarily be able to simulate a quantum computer.
Thus, any general method for simulating quantum mechanics with at most
a polynomial slowdown would lead to a polynomial-time algorithm for
factoring.
The rest of this paper is organized as follows. In \S2,
we introduce the model of quantum computation, the {\em quantum gate array,}
that we use in the rest of the paper. In \S\S3 and~4,
we explain two subroutines that are used in our algorithms:
reversible modular exponentiation in~\S3
and quantum Fourier transforms in~\S4. In \S5, we give our
algorithm for prime factorization, and
in \S6, we give our algorithm for extracting discrete
logarithms. In \S7, we give a brief discussion of the
practicality of quantum computation and suggest possible directions for
further work.
\section{Quantum computation}
In this section we give a brief introduction to quantum computation,
emphasizing the properties that we will use.
We will describe only {\em quantum gate arrays,} or {\em quantum
acyclic circuits,} which are analogous to acyclic circuits in
classical computer science. For other models
of quantum computers, see references
on quantum Turing machines \cite{Deut89,BeVa,Yao} and quantum cellular
automata [Feynman 1986, Margolus 1986, 1990, Lloyd 1993, Biafore 1994].
If they are allowed a small probability of
error, quantum Turing machines and quantum gate arrays can compute the same
functions in polynomial time \cite{Yao}. This may also be true for the
various models
of quantum cellular automata, but it has not yet been proved.
This gives evidence that the class of functions computable in quantum
polynomial time with a small probability of error is robust, in that it
does not depend on the exact architecture of a quantum computer. By analogy
with the classical class BPP, this class is called BQP.
Consider a system with $n$ components, each of which can have two states.
Whereas in classical physics, a complete description of the
state of this system requires
only $n$ bits, in quantum physics, a complete description of the state of
this system requires $2^n-1$ complex numbers. To be more
precise, the state of the quantum
system is a point in a $2^n$-dimensional vector space.
For each of the $2^n$ possible classical
positions of the components, there is a basis state
of this vector space which we represent, for example,
by $\|0 1 1 \cdots 0\>$ meaning that the first bit is~0, the second
bit is~1, and so on. Here, the {\em ket} notation $\|x\>$ means that
$x$ is a (pure) quantum state. (Mixed states will not be discussed in
this paper, and thus we do not define them; see a quantum theory book
such as Peres [1993] for this definition.)
The {\em Hilbert space} associated with this quantum system is
the complex vector space with these $2^n$ states as basis vectors, and
the state of the system at any time is represented by a unit-length vector
in this Hilbert space.
As multiplying this state vector by a unit-length complex phase does
not change any behavior of the state, we need only $2^n-1$ complex
numbers to completely describe the state.
We represent this superposition of states as
\begin{equation}
\sum_{i=0}^{2^n-1} a_i \| S_i \>,
\end{equation}
where the amplitudes $a_i$ are complex numbers such that
$\sum_i \left| a_i \right| ^2 =1$ and
each $\| S_i \>$ is a basis vector of the Hilbert space.
If the machine is measured (with respect to this basis)
at any particular step, the probability of
seeing basis state $\| S_i \>$ is $\left| a_i \right| ^2$; however,
measuring the state of the machine projects this state to the
observed basis vector $\|S_i\>$.
Thus, looking at the machine during the computation will invalidate the rest
of the computation. In this paper, we only consider measurements with
respect to the canonical basis. This does not greatly restrict our
model of computation, since measurements in other reasonable
bases could be simulated by first using quantum computation to
perform a change of basis and then performing a measurement in the
canonical basis.
In order to use a physical
system for computation, we must be able to change the state of the system.
The laws of quantum mechanics permit only unitary transformations of
state vectors. A unitary matrix is one whose conjugate transpose is
equal to its inverse, and requiring state transformations to be represented
by unitary matrices ensures that summing the probabilities
of obtaining every possible outcome will result in~1.
The definition of quantum circuits (and quantum Turing machines) only
allows {\em local} unitary transformations; that is, unitary transformations
on a fixed number of bits.
This is physically justified because,
given a general unitary transformation on $n$ bits, it is
not at all clear how one would efficiently implement it physically, whereas
two-bit transformations can at least in theory be implemented by relatively
simple physical systems \cite{CiZo,DiVi,SlWe,ChYa}. While general $n$-bit
transformations can always be built out of two-bit transformations
[DiVincenzo 1995, Sleator and Weinfurter 1995, Lloyd 1995, Deutsch et al.\
1995], the number required will often be
exponential in $n$ \cite{nine}. Thus, the set of two-bit transformations
form a set of building blocks for quantum circuits in a manner analogous
to the way a universal set of classical gates (such as the
AND, OR and NOT gates) form a set of building blocks for classical circuits.
In fact, for a universal set of quantum gates, it is sufficient to take
all one-bit gates and a single type of two-bit gate, the controlled NOT,
which negates the second bit if and only if the first bit is~1.
Perhaps an example will be informative at this point. A quantum gate
can be expressed as a truth table: for each input basis vector we need
to give the output of the gate. One such gate is:\\
\noindent
\begin{minipage}{\textwidth}
\begin{eqnarray}
\|00\> &\rightarrow& \|00\> \nonumber \\
\|01\> &\rightarrow& \|01\> \label{exampletrtable} \\
\|10\> &\rightarrow&
{\textstyle \frac{1}{\sqrt{2}}} (\|10\> + \|11\>) \nonumber \\
\|11\> &\rightarrow&
{\textstyle \frac{1}{\sqrt{2}}} (\|10\> - \|11\>). \nonumber
\end{eqnarray}
\end{minipage}
\vspace{\baselineskip}
\noindent
Not all truth tables correspond to physically feasible quantum gates,
as many truth tables will not give rise to unitary transformations.
The same gate can also be represented as a matrix. The rows correspond
to input basis vectors. The columns correspond to output basis vectors.
The $(i,j)$ entry gives, when the $i$th basis vector is input to the gate,
the coefficient of the $j$th basis vector in the corresponding output of
the gate. The truth table above would then correspond to the following
matrix:
\begin{equation}
\begin{array}{c|cccc|l}
\multicolumn{1}{c}{} & \|00\> & \|01\> & \|10\>
& \multicolumn{1}{c}{\|11\>} & \\*[.5ex]
\|00\> \ \ & 1 & 0 & 0 & 0 & \\*[.5ex]
\|01\> \ \ & 0 & 1 & 0 & 0 & \\*[.5ex]
\|10\> \ \ & 0 & 0 & \frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} & \\*[.5ex]
\|11\> \ \ & 0 & 0 & \frac{1}{\sqrt{2}} & -\frac{1}{\sqrt{2}}\phantom{-} &.
\end{array}
\label{examplematrix}
\end{equation}
A quantum gate is feasible if and only if the corresponding matrix is
unitary, i.e., its inverse is its conjugate transpose.
Suppose our
machine is in the superposition of states
\begin{equation}
\textstyle
\frac{1}{\sqrt{2}} \|10\> - \frac{1}{\sqrt{2}} \|11\>
\label{examplevector}
\end{equation}
and we apply the unitary transformation represented
by (\ref{exampletrtable}) and
(\ref{examplematrix}) to this state. The resulting output will be the
result of multiplying the vector (\ref{examplevector}) by
the matrix (\ref{examplematrix}). The machine will thus go to the
superposition of states
\begin{equation}
\textstyle
\frac{1}{2}\left(\|10\> + \|11\>\right)
-
\frac{1}{2}\left(\|10\> - \|11\>\right) \;=\;
\|11\>.
\end{equation}
This example shows the potential effects of interference on quantum
computation. Had we started with either the state $\|10\>$ or the
state $\|11\>$, there would have been a chance of observing the
state $\|10\>$ after the application of the gate (\ref{examplematrix}).
However, when we start with a superposition of these two states, the
probability amplitudes for the state $\|10\>$ cancel, and we have no
possibility of observing $\|10\>$ after the application of the gate.
Notice that the output of the
gate would have been $\|10\>$ instead of $\|11\>$ had
we started with the superposition of states
\begin{equation}
\textstyle
\frac{1}{\sqrt{2}} \|10\> + \frac{1}{\sqrt{2}} \|11\>
\end{equation}
which has the same probabilities of being in any particular configuration
if it is observed as does the superposition (\ref{examplevector}).
If we apply a gate to only two bits of a longer basis vector (now
our circuit must have more than two wires), we multiply the
gate matrix by
the two bits to which the gate is applied, and leave the other bits alone.
This corresponds to multiplying the whole state by the tensor product
of the gate matrix on those two bits with the
identity matrix on the remaining bits.
A quantum gate array is a set of quantum gates with
logical ``wires'' connecting
their inputs and outputs. The input to the
gate array, possibly along with extra work bits that
are initially set to~0, is fed through a sequence of quantum gates.
The values of the bits are observed after the last quantum gate,
and these values are the output. To compare
gate arrays with quantum Turing machines, we need to add conditions
that make gate arrays a {\em uniform} complexity class. In other words,
because there is a different gate array for each size of input, we
need to keep the designer of the gate arrays from hiding non-computable
(or hard to compute) information in the arrangement of the gates.
To make quantum gate arrays uniform, we must add two things to the
definition of gate arrays. The first is the standard requirement that
the design of the gate array be produced by a polynomial-time (classical)
computation. The second requirement
should be a standard part of the definition of analog complexity classes,
although since analog complexity classes have not
been widely studied, this requirement is much less widely known.
This requirement is that the entries in the unitary matrices describing
the gates must be computable numbers. Specifically, the first $\log n$
bits of each entry should be classically computable in time polynomial in~$n$
\cite{Solo}.
This keeps non-computable (or hard to compute) information
from being hidden in the bits of the amplitudes of the quantum gates.
\section{Reversible logic and modular exponentiation}
The definition of quantum gate arrays gives rise to completely reversible
computation. That is, knowing the quantum state on the wires leading out
of a gate tells uniquely what the quantum state must have been on the
wires leading into that gate. This is a reflection of the fact that,
despite the macroscopic arrow of time, the laws of physics appear to
be completely reversible. This would seem to imply that anything built
with the laws
of physics must be completely reversible; however, classical computers
get around this fact by dissipating energy and thus making their
computations thermodynamically irreversible. This appears
impossible to do for quantum computers because superpositions of
quantum states need to be maintained throughout the computation. Thus,
quantum computers necessarily have to use reversible computation.
This imposes extra costs when doing classical computations on a quantum
computer, as is sometimes necessary in subroutines of quantum
computations.
Because of the reversibility of quantum computation,
a deterministic computation is performable on a quantum
computer only if it is reversible.
Luckily, it has already been shown that any deterministic computation
can be made reversible \cite{Lece,Benn73}. In fact, reversible
classical gate arrays have been studied. Much like the result that any
classical computation can be done using NAND gates, there are also universal
gates for reversible computation. Two of these are Toffoli gates
\cite{Toff} and Fredkin gates \cite{FrTo}; these are illustrated in
Table~\ref{revgates}.
\begin{table}
\caption{Truth tables for Toffoli and Fredkin gates.}
\footnotesize
\hspace*{\fill}
\begin{tabular}%
{|@{\hspace{1em}}ccc@{\hspace{1em}}|@{\hspace{1em}}ccc@{\hspace{1em}}|}
\multicolumn{6}{c}{Toffoli Gate}\\
\multicolumn{6}{c}{}\\
\hline
\multicolumn{3}{|c|@{\hspace{1em}}}{INPUT}&
\multicolumn{3}{@{\hspace{-1em}}c@{\hspace{0em}}|}{OUTPUT}\\
\hline
0 & 0 & 0 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 & 0 & 1 \\
0 & 1 & 0 & 0 & 1 & 0 \\
0 & 1 & 1 & 0 & 1 & 1 \\
1 & 0 & 0 & 1 & 0 & 0 \\
1 & 0 & 1 & 1 & 0 & 1 \\
1 & 1 & 0 & 1 & 1 & 1 \\
1 & 1 & 1 & 1 & 1 & 0 \\
\hline
\end{tabular}
\hspace*{\fill}
\begin{tabular}%
{|@{\hspace{1em}}ccc@{\hspace{1em}}|@{\hspace{1em}}ccc@{\hspace{1em}}|}
\multicolumn{6}{c}{Fredkin Gate}\\
\multicolumn{6}{c}{}\\
\hline
\multicolumn{3}{|c|@{\hspace{1em}}}{INPUT}&
\multicolumn{3}{@{\hspace{-1em}}c@{\hspace{0em}}|}{OUTPUT}\\
\hline
0 & 0 & 0 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 & 1 & 0 \\
0 & 1 & 0 & 0 & 0 & 1 \\
0 & 1 & 1 & 0 & 1 & 1 \\
1 & 0 & 0 & 1 & 0 & 0 \\
1 & 0 & 1 & 1 & 0 & 1 \\
1 & 1 & 0 & 1 & 1 & 0 \\
1 & 1 & 1 & 1 & 1 & 1 \\
\hline
\end{tabular}
\hspace*{\fill}
\label{revgates}
\end{table}
The Toffoli gate is just a controlled controlled NOT, i.e., the
last bit is negated if and only if the first two bits are~1.
In a Toffoli gate, if the third input bit is set to~1, then the
third output bit is the NAND of the first two input bits. Since
NAND is a universal gate for classical gate arrays, this shows
that the Toffoli gate is universal. In a Fredkin gate, the
last two bits are swapped if the first bit is~0, and left untouched
if the first bit is~1. For a Fredkin gate, if the third input bit
is set to~0, the second output bit is the AND of the first two input
bits; and if the last two input bits are set to 0 and 1 respectively,
the second output bit is the NOT of the first input bit. Thus, both
AND and NOT gates are realizable using Fredkin gates, showing that
the Fredkin gate is universal.
From results on
reversible computation \cite{Lece,Benn73}, we can compute
any polynomial time function $F(x)$ as long as we keep the input $x$
in the computer. We do this by adapting the method for computing the
function~$F$ non-reversibly. These results can easily be extended
to work for gate arrays \cite{Toff,FrTo}.
When AND, OR or NOT gates are changed to Fredkin or Toffoli gates,
one obtains both additional input bits, which must be preset to specified
values, and additional output bits, which contain the information needed to
reverse the computation. While the additional input bits
do not present difficulties in designing quantum computers, the
additional output bits do, because unless they are all reset to~0,
they will affect the interference patterns in quantum computation.
Bennett's method for resetting these bits to 0 is shown in the top half
of Table~\ref{reverse}. A non-reversible gate array may thus be turned
into a reversible gate array as follows. First, duplicate the
input bits as many times as necessary (since each input bit could be used
more than once by the gate array). Next, keeping one copy of the input
around, use Toffoli and Fredkin gates to simulate non-reversible gates,
putting the extra output bits into the RECORD register. These extra output
bits preserve enough of a record of the operations to enable the computation
of the gate array to be reversed.
Once the output $F(x)$ has been computed, copy it into a register that
has been preset to zero, and then undo the computation to erase both
the first OUTPUT register and the RECORD register.
\begin{table}[ht]
\caption{Bennett's method for making a computation reversible.}
\footnotesize
\begin{center}
\begin{tabular}
{|@{\hspace{2em}}p{7.5em}p{7.5em}p{7.5em}p{7.5em}@{\hspace{-1em}}|}
\hline
INPUT & - - - - - - & - - - - - - & - - - - - - \\
INPUT & OUTPUT & RECORD($F$) & - - - - - - \\
INPUT & OUTPUT & RECORD($F$) & OUTPUT \\
INPUT & - - - - - - & - - - - - - & OUTPUT \\
\hline
INPUT & INPUT & RECORD($F^{-1}$) & OUTPUT \\
- - - - - - & INPUT & RECORD($F^{-1}$) & OUTPUT \\
- - - - - - & - - - - - - & - - - - - - & OUTPUT\\
\hline
\end{tabular}
\hspace{3em}
\end{center}
\label{reverse}
\end{table}
To erase $x$ and replace it with $F(x)$, in addition to a
polynomial-time algorithm for~$F$, we also need
a polynomial-time algorithm for computing $x$
from $F(x)$; i.e., we need that $F$ is one-to-one and
that both $F$ and $F^{-1}$ are polynomial-time computable.
The method for this computation
is given in the whole of Table~\ref{reverse}.
There are two stages to this computation. The first is the same as
before, taking $x$ to $(x,F(x))$. For the second stage, shown in the
bottom half of Table~\ref{reverse}, note that if
we have a method to compute $F^{-1}$ non-reversibly in polynomial time,
we can use the same technique to reversibly
map $F(x)$ to $(F(x), F^{-1}(F(x))) = (F(x),x)$. However, since this
is a reversible computation, we can reverse it
to go from $(x,F(x))$ to $F(x)$. Put together,
these two pieces take $x$ to $F(x)$.
The above discussion shows that computations can be made reversible for
only a constant factor cost in time, but the above method uses as much space
as it does time. If the classical computation requires much less space
than time, then making it reversible in this manner will result
in a large increase
in the space required. There are methods that do not use as much
space, but use more time, to
make computations reversible \cite{Benn89,LeSh}. While there is no
general method that does not cause an increase in either space or time,
specific algorithms can sometimes be made reversible without paying
a large penalty in either space or time; at the end of this
section we will show how to do
this for modular exponentiation, which is a subroutine necessary for
quantum factoring.
The bottleneck in the quantum factoring algorithm; i.e., the piece
of the factoring algorithm that consumes the most time and
space, is modular exponentiation. The modular exponentiation
problem is, given $n$, $x$, and $r$, find $x^r \mod{n}$.
The best classical method for doing this is to repeatedly square of
$x\mod{n}$ to get $x^{2^i}\mod{n}$ for $i\leq\log_2 r$, and then multiply a
subset of these powers $\bmod{n}$ to get $x^r\mod{n}$. If
we are working with $l$-bit numbers, this requires $O(l)$ squarings
and multiplications of $l$-bit numbers $\bmod{n}$. Asymptotically,
the best classical result for gate arrays for
multiplication is the Sch\"onhage--Strassen algorithm \cite{ScSt,Knut,Scho}.
This gives a gate array for integer multiplication that uses
$O(l \log l \log \log l)$ gates to multiply two $l$-bit numbers. Thus,
asymptotically, modular exponentiation requires $O(l^2 \log l \log \log l)$
time. Making this reversible would na\"\i vely cost the same amount in
space; however, one can reuse the space used in the repeated squaring part
of the algorithm, and thus reduce the amount of space needed to essentially
that required for multiplying two $l$-bit numbers; one simple
method for reducing this space (although
not the most versatile one) will be given later in this
section. Thus, modular exponentiation can be done in
$O(l^2 \log l \log \log l)$ time and $O(l \log l \log \log l )$ space.
While the Sch\"onhage--Strassen algorithm is the best
multiplication algorithm discovered
to date for large~$l$, it does not scale well for small~$l$.
For small numbers, the best gate arrays for multiplication essentially use
elementary-school longhand multiplication in binary. This method requires
$O(l^2)$ time to multiply two $l$-bit numbers, and thus modular
exponentiation requires $O(l^3)$ time with this method. These gate arrays
can be made reversible, however, using only $O(l)$ space.
We will now give the method for constructing a reversible gate array that
takes only $O(l)$ space and $O(l^3)$ time to
compute $(a,x^a\mod{n})$ from~$a$, where $a$, $x$, and $n$
are $l$-bit numbers. The basic building
block used is a gate array that takes $b$ as input and outputs
$b+c\mod{n}$. Note that here $b$ is the gate array's input but $c$ and $n$
are built into the structure of the gate array.
Since addition $\bmod{n}$ is
computable in $O(\log n)$ time classically, this reversible gate array
can be made with only $O(\log n)$ gates and $O(\log n)$ work
bits using the techniques explained earlier in this section.
The technique we use for computing $x^a\mod{n}$ is essentially the same
as the classical method. First, by repeated squaring we compute
$x^{2^i}\mod{n}$ for all $i < l$. Then, to obtain $x^a\mod{n}$ we
multiply the powers $x^{2^i}$ $\bmod{n}$
where $2^i$ appears in the binary expansion of~$a$.
In our algorithm for factoring~$n$, we only
need to compute $x^{a}\mod{n}$ where $a$ is in a superposition of states,
but $x$ is some fixed integer. This makes things much easier,
because we can use a reversible gate array where $a$ is
treated as input, but where $x$ and $n$ are built
into the structure of the gate array. Thus, we can use the algorithm
described by the following pseudocode;
here, $a_i$ represents the $i$th bit of $a$ in binary,
where the bits are indexed from right to left and the rightmost
bit of $a$ is~$a_0$.
\vspace{\baselineskip}
\noindent
\begin{minipage}{\textwidth}
{\tt
\begin{tabbing}
\ \ \ \ \ \ \ \ \ \= \ \ \ \ \= \ \ \ \ \= \kill
\> {\it power} := 1 \\
\> for {\it i} = 0 to {\it l}$\,-$1 \\
\> \> if ( {\it a}$_i$ $==$ 1 ) then \\
\> \> \> {\it power} := {\it power} $*$ {\it x}$^{\,2^i}$ $\mod{n}$ \\
\> \> endif\\
\> endfor
\end{tabbing}
}
\end{minipage}
\vspace{\baselineskip}
\noindent
The variable $a$ is left unchanged by the code and $x^a\mod{n}$
is output as the variable ${\it power}$. Thus, this code takes
the pair of values $(a,1)$ to $(a,x^a\mod{n})$.
This pseudocode can easily be turned into a gate array;
the only hard part of this is the fourth line, where we multiply the
variable {\it power} by $x^{2^i}$ $\bmod{n}$; to do this we need to use a
fairly complicated gate array as a subroutine.
Recall that $x^{2^i}\mod{n}$ can be computed classically and then
built into the
structure of the gate array. Thus, to implement this line, we need
a reversible gate array that takes $b$ as input and gives $bc \mod{n}$ as
output, where the structure of the gate array can depend on $c$ and~$n$.
Of course, this step
can only be reversible if $\gcd(c,n) =1$, i.e., if
$c$ and $n$ have
no common factors, as otherwise two distinct values of $b$ will be
mapped to the same value of $bc \mod{n}$; this case is
fortunately all we need
for the factoring algorithm. We will show how
to build this gate array in two stages. The first stage is directly
analogous to exponentiation by repeated multiplication; we obtain
multiplication from repeated addition $\bmod{n}$. Pseudocode for this
stage is as follows.
\vspace{\baselineskip}
\noindent
\begin{minipage}{\textwidth}
{\tt
\begin{tabbing}
\ \ \ \ \ \ \ \ \ \= \ \ \ \ \= \ \ \ \ \= \kill
\> {\it result} := 0\\
\> for {\it i} = 0 to {\it l}$\,-$1\\
\> \> if ( {\it b}$_i$ $==$ 1 ) then \\
\> \> \> {\it result} := {\it result} $+$ 2$^{i}${\it c} $\mod{n}$\\
\> \> endif\\
\> endfor
\end{tabbing}
}
\end{minipage}
\vspace{\baselineskip}
\noindent
Again, $2^ic\mod{n}$ can be precomputed and built into the
structure of the gate array.
The above pseudocode takes $b$ as input, and gives $(b,bc\mod{n})$ as
output. To get the desired result, we now
need to erase~$b$. Recall that $\gcd(c,n)=1$, so
there is a $c^{-1}\mod{n}$ with $c \, c^{-1} \equiv 1\mod{n}$.
Multiplication by this $c^{-1}$ could be used to reversibly take
$bc\mod{n}$ to $(bc\mod{n},bcc^{-1}\mod{n}) = (bc\mod{n},b)$. This
is just the reverse of the operation we want, and since we are working
with reversible computing, we can turn this operation around to erase~$b$.
The pseudocode for this follows.
\vspace{\baselineskip}
\noindent
\begin{minipage}{\textwidth}
{\tt
\begin{tabbing}
\ \ \ \ \ \ \ \ \ \= \ \ \ \ \= \ \ \ \ \= \kill
\> for {\it i} = 0 to {\it l}$\,-$1\\
\> \> if ( {\it result}$_{\,i}$ $==$ 1 ) then\\
\> \> \> {\it b} := {\it b} $-$ 2$^{i}${\it c}$^{-1}$ $\mod{n}$\\
\> \> endif\\
\> endfor
\end{tabbing}
}
\end{minipage}
\vspace{\baselineskip}
\noindent
As before, {\it result}$_{\,i}$ is the $i$th bit of {\it result}.
Note that at this stage of the computation, $b$ should be~0.
However, we did not set
$b$ directly to zero, as this would not have been a reversible
operation and thus impossible on a quantum computer,
but instead we did a relatively complicated sequence
of operations which ended with $b=0$ and which in fact depended on
multiplication being a group $\bmod{n}$. At
this point, then, we could do something somewhat sneaky: we could
measure $b$ to see if it actually is~0. If it is not, we know that
there has been an error somewhere in the quantum computation,
i.e., that the results are worthless and we should stop the computer
and start over again. However, if we do find that $b$ is~0, then we know
(because we just observed it) that it is now exactly~0. This
measurement thus may bring the quantum computation back on track
in that any amplitude that $b$ had for being non-zero has been
eliminated. Further, because the probability that we observe a state
is proportional to the square of the amplitude of that state, depending
on the error model, doing the
modular exponentiation and measuring $b$ every time that we know
that it should be 0 may have a higher probability of overall success
than the same computation done without the repeated measurements of~$b$;
this is the {\em quantum watchdog} (or {\em quantum Zeno}) effect
\cite{Pere2}.
The argument above does not actually show that repeated measurement of
$b$ is indeed beneficial, because there is a cost (in time, if nothing
else) of measuring~$b$. Before this is implemented, then, it should
be checked with analysis or experiment that the benefit of
such measurements exceeds their cost. However, I believe that
partial measurements such as this one are a promising way of trying
to stabilize quantum computations.
Currently, Sch\"onhage--Strassen is the algorithm of choice for
multiplying very large numbers, and longhand multiplication is the algorithm
of choice for small numbers.
There are also multiplication algorithms which have efficiencies between
these two algorithms, and which are the best
algorithms to use for intermediate length numbers \cite{KaOf,Knut,ScGrVe}.
It is not clear which algorithms are best for which size numbers.
While this may be known to some extent for classical computation
\cite{ScGrVe}, using data on which algorithms work better on classical
computers could be misleading for two reasons: First, classical computers
need not be reversible, and the cost of making an algorithm reversible
depends on the algorithm. Second, existing computers generally have
multiplication for 32- or 64-bit numbers built into their hardware,
and this will increase the optimal changeover points to asymptotically
faster algorithms; further, some multiplication algorithms can take
better advantage of this hardwired multiplication than others. Thus,
in order to program quantum computers most efficiently, work needs
to be done on the best way of implementing elementary arithmetic
operations on quantum computers. One tantalizing fact is that the
Sch\"onhage--Strassen fast multiplication algorithm uses the fast
Fourier transform, which is also the basis for all the fast algorithms
on quantum computers discovered to date; it is tempting to speculate
that integer multiplication itself might be speeded up by a quantum
algorithm; if possible, this would result in a somewhat faster asymptotic
bound for factoring on a quantum computer, and indeed could even
make breaking RSA on a quantum computer asymptotically faster
than encrypting with RSA on a classical computer.
\section{Quantum Fourier transforms}
Since quantum computation deals with unitary transformations, it is helpful
to be able to build certain useful unitary transformations. In
this section we give a technique for constructing in polynomial time
on quantum computers one particular unitary transformation,
which is essentially a discrete Fourier transform. This transformation
will be given as a matrix, with both rows and columns indexed
by states. These states correspond to binary
representations of integers on the computer; in particular, the rows
and columns will be indexed beginning with 0 unless otherwise specified.
This transformations is as follows.
Consider a number $a$ with $0 \leq a < q$ for some
$q$ where the number of bits of $q$ is polynomial. We will perform the
transformation that takes the state $\|a\>$ to the state
\begin{equation}
\frac{1}{q^{1/2}} \sum_{c=0}^{q-1} \|c\> \exp(2 \pi i ac/q).
\label{ft}
\end{equation}
That is, we apply
the unitary matrix whose
$(a,c)$ entry is
$\frac{1}{q^{1/2}}\exp(2 \pi i ac/q)$.
This Fourier transform is at the heart of our algorithms, and we call
this matrix~$A_q$.
Since we will use $A_q$ for $q$ of exponential size, we must show how this
transformation can be done in polynomial time. In this paper, we will
give a simple construction for $A_q$ when $q$ is a power of~2 that was
discovered independently by Coppersmith [1994] and Deutsch
[see Ekert and Jozsa 1995]. This construction is essentially the
standard fast Fourier transform (FFT) algorithm \cite{Knut}
adapted for a quantum computer; the following description of it
follows that of Ekert and Jozsa [1995].
In the earlier version of this paper [Shor 1994], we gave a
construction for $A_q$ when $q$ was in the special class of smooth numbers
with small prime power factors. In fact, Cleve [1994] has shown how
to construct $A_q$ for all smooth numbers $q$
whose prime factors are at most $O(\log n)$.
Take $q=2^l$, and let us represent an integer $a$ in binary as
$\|a_{l-1}a_{l-2}\ldots a_0\>$.
For the quantum Fourier transform~$A_q$, we only need to use two types of
quantum gates. These gates are $R_j$, which operates on the $j$th
bit of the quantum computer:
\begin{equation}
R_j \ = \
\begin{array}{c|cc|l}
\multicolumn{1}{c}{} & \|0\> & \multicolumn{1}{c}{\|1\>} & \\*[.5ex]
\|0\> \ \ & \frac{1}{\sqrt{2}} & \phantom{-} \frac{1}{\sqrt{2}} & \\*[.5ex]
\|1\> \ \ & \frac{1}{\sqrt{2}} & -\frac{1}{\sqrt{2}}& ,
\end{array}
\label{Rmatrix}
\end{equation}
and
$S_{j,k}$, which operates on the bits in positions $j$ and $k$ with $j < k$:
\begin{equation}
S_{j,k} \ = \
\begin{array}{c|cccc|l}
\multicolumn{1}{c}{} & \|00\> & \|01\> & \|10\> &
\multicolumn{1}{c}{\|11\>} & \\*[.5ex]
\|00\> \ \ & 1 & 0 & 0 & 0 & \\*[.5ex]
\|01\> \ \ & 0 & 1 & 0 & 0 & \\*[.5ex]
\|10\> \ \ & 0 & 0 & 1 & 0 & \\*[.5ex]
\|11\> \ \ & 0 & 0 & 0 & e^{i\theta_{k-j}} & ,
\end{array}
\label{Smatrix}
\end{equation}
where $\theta_{k-j} = \pi/2^{k-j}$.
To perform a quantum Fourier transform, we apply the matrices in the
order (from left to right)
\begin{equation}
\, R_{l-1}\, S_{l-2,l-1}\, R_{l-2}\, S_{l-3,l-1}\, S_{l-3,l-2}\, R_{l-3}
\ldots R_1 \,S_{0,l-1}\, S_{0,l-2} \ldots S_{0,2}\, S_{0,1}\, R_0\, ;
\end{equation}
that is, we apply the gates $R_j$ in reverse order from $R_{l-1}$ to~$R_0$,
and between $R_{j+1}$ and $R_j$ we apply all the gates $S_{j,k}$ where $k>j$.
For example, on 3 bits, the matrices would be applied in the order
$R_2 S_{1,2} R_1 S_{0,2} S_{0,1} R_0$.
To take the Fourier transform
$A_q$ when $q=2^l$, we thus need to use $l(l-1)/2$ quantum gates.
Applying this sequence of transformations will result in a quantum state
$\frac{1}{q^{1/2}} \sum_b \exp(2 \pi i ac /q) \|b\>$,
where $b$ is the bit-reversal of~$c$,
i.e., the binary number obtained by reading the bits of $c$ from right
to left. Thus, to obtain the actual quantum Fourier transform, we need
either to do further computation to
reverse the bits of $\|b\>$ to obtain $\|c\>$, or to leave these bits
in place and read them in reverse order; either alternative is easy
to implement.
To show that this operation actually performs a quantum Fourier transform,
consider the amplitude of going from $\|a\> = \|a_{l-1} \ldots a_0\>$
to $\|b\> = \|b_{l-1} \ldots b_0\>$. First, the factors of $1/\sqrt{2}$
in the $R$ matrices multiply to produce a factor of $1/q^{1/2}$ overall;
thus we need only worry about the $\exp(2 \pi i ac/q)$ phase
factor in the
expression (\ref{ft}). The matrices $S_{j,k}$ do not change the values of
any bits, but merely change their phases. There is thus only one way to
switch the $j$th bit from $a_j$ to~$b_j$, and that is to use the appropriate
entry in the matrix~$R_j$. This entry adds $\pi$ to the phase if the
bits $a_j$ and $b_j$ are both~1, and leaves it unchanged otherwise.
Further, the matrix $S_{j,k}$ adds $\pi/2^{k-j}$
to the phase if $a_j$ and $b_k$ are both 1 and leaves it unchanged otherwise.
Thus, the phase on the path from $\|a\>$ to $\|b\>$ is
\begin{equation}
\sum_{0 \leq j < l} \pi a_jb_j +
\sum_{0\leq j<k < l} \frac{\pi}{2^{k-j}} a_j b_k.
\end{equation}
This expression can be rewritten as
\begin{equation}
\sum_{0 \leq j \leq k < l} \frac{\pi}{2^{k-j}} a_j b_k.
\end{equation}
Since $c$ is the bit-reversal of~$b$, this expression can be
further rewritten as
\begin{equation}
\sum_{0 \leq j \leq k < l} \frac{\pi}{2^{k-j}} a_j c_{l-1-k}.
\end{equation}
Making the substitution $l-k-1$ for $k$ in this sum, we get
\begin{equation}
\sum_{0 \leq j+k < l} 2 \pi \frac{2^j 2^k}{2^l} a_j c_{k}
\end{equation}
Now, since adding multiples of $2 \pi$ do not affect the phase, we obtain
the same phase if we sum over all $j$ and $k$ less than~$l$, obtaining
\begin{equation}
\sum_{j,k =0}^{l-1} 2 \pi \frac{2^j 2^k}{2^l} a_j c_{k} = \frac{2\pi}{2^l}\;
\sum_{j=0}^{l-1} 2^j a_j \;\sum_{k=0}^{l-1} 2^k c_k,
\end{equation}
where the last equality follows
from the distributive law of multiplication.
Now, $q=2^l$, $a = \sum_{j=0}^{l-1}2^j a_j$, and similarly for~$c$, so
the above expression is equal to
$2 \pi ac/q$,
which is the phase for the amplitude of $\|a\> \rightarrow \|c\>$ in
the transformation (\ref{ft}).
When $k-j$ is large in the gate $S_{j,k}$ in (\ref{Smatrix}), we are
multiplying by a
very small phase factor. This would be very difficult to do
accurately physically, and thus it would be somewhat disturbing if this
were necessary for quantum computation. Luckily, Coppersmith [1994]
has shown that one can define an approximate Fourier transform that ignores
these tiny phase factors, but which approximates the Fourier transform
closely enough that it can also be used for factoring. In fact,
this technique reduces the number of quantum gates needed for the
(approximate) Fourier transform considerably, as it leaves out most of
the gates~$S_{j,k}$.
\section{Prime factorization}
It has been known since before Euclid that every integer $n$ is uniquely
decomposable into a product of primes. Mathematicians have been
interested in the question of how to factor a number into this
product of primes for nearly as long. It was only in the 1970's,
however, that researchers applied the paradigms of theoretical computer
science to number theory, and looked at the asymptotic running times
of factoring algorithms \cite{Adle}. This has resulted in a great
improvement in the efficiency of factoring algorithms. The best
factoring algorithm asymptotically is currently the number field sieve
\cite{LeLeMaPo,LeLe}, which in order to factor an integer $n$ takes
asymptotic running time
$\exp(c(\log n)^{1/3} (\log \log n)^{2/3})$ for some constant~$c$.
Since the input, $n$, is only $\log n$ bits in length, this
algorithm is an exponential-time algorithm.
Our quantum factoring algorithm takes asymptotically
$O((\log n)^2 \linebreak[1] (\log \log n) \linebreak[1] (\log \log \log n))$
steps on a quantum computer, along with a polynomial (in $\log n$) amount
of post-processing time on a classical computer that is
used to convert the output of the quantum computer to factors
of~$n$. While this post-processing could in principle be done on a quantum
computer, there is no reason not to use a classical computer if they are
more efficient in practice.
Instead of giving a quantum computer algorithm for factoring $n$ directly,
we give a quantum computer algorithm for finding the order of
an element $x$ in the multiplicative group
$\bmod{n}$; that is, the least integer $r$ such that
$x^r \equiv 1 \mod{n}$. It is known that using randomization,
factorization can be reduced to finding the order of an element \cite{Mill};
we now briefly give this reduction.
To find a factor of an odd number~$n$, given a method for computing
the order $r$ of~$x$, choose a random $x \mod{n}$,
find its order~$r$, and compute $\gcd(x^{r/2}-1,n)$. Here,
$\gcd(a,b)$ is the greatest common divisor of $a$ and~$b$, i.e.,
the largest integer that divides both $a$ and~$b$. The
Euclidean algorithm \cite{Knut} can be used to compute $\gcd(a,b)$
in polynomial time. Since
$(x^{r/2}-1)\linebreak[1](x^{r/2}+1) = x^{r}-1 \equiv 0 \mod{n}$,
the $\gcd(x^{r/2}-1,n)$ fails to be
a non-trivial divisor of $n$ only if $r$ is odd or if $x^{r/2}
\equiv -1 \mod{n}$. Using this criterion, it can be shown that this
procedure, when applied to a random~$x \mod{n}$,
yields a factor of $n$ with probability
at least $1-1/2^{k-1}$, where $k$
is the number of distinct odd prime factors of~$n$. A brief sketch of the
proof of this result follows. Suppose that $n = \prod_{i=1}^{k} p_i^{a_i}$.
Let $r_i$ be the order of $x \mod{p_i^{a_i}}$.
Then $r$ is the least common multiple of all the~$r_i$.
Consider the largest power of 2 dividing each~$r_i$. The algorithm
only fails if all of these powers of 2 agree: if they are all~1, then
$r$ is odd and $r/2$ does not exist; if they are all equal and
larger than~1, then
$x^{r/2} \equiv -1 \mod{n}$ since $x^{r/2} \equiv -1 \mod{p_i^{\alpha_i}}$
for every~$i$. By the Chinese remainder
theorem [Knuth 1981, Hardy and Wright 1979, Theorem 121], choosing an
$x \mod{n}$ at random is the same as choosing
for each $i$ a number $x_i \mod{p_i^{a_i}}$ at random, where $p_i^{a_i}$
is the $i$th prime power
factor of~$n$. The multiplicative group $\bmod{p^{\alpha}}$ for
any odd prime power $p^{\alpha}$ is
cyclic \cite{Knut}, so for any odd prime power $p_i^{a_i}$, the
probability is at most $1/2$ of
choosing an $x_i$ having any particular power of two as the largest divisor
of its order~$r_i$.
Thus each of these powers of 2 has at most a 50\% probability
of agreeing with the previous ones, so all $k$ of them agree with
probability at most $1/2^{k-1}$,
and there is at least a $1-1/2^{k-1}$ chance that the $x$ we choose
is good.
This scheme will thus
work as long as $n$ is odd and not a prime power; finding factors of
prime powers can be done efficiently with classical methods.
We now describe the algorithm for finding the
order of $x \mod{n}$ on a quantum computer.
This algorithm will use two quantum registers which hold integers represented
in binary. There will also be some amount of workspace. This workspace
gets reset to 0 after each subroutine of our algorithm, so we will
not include it when we write down the state of our machine.
Given $x$ and~$n$, to find the order of~$x$, i.e., the least $r$ such
that $x^r \equiv 1 \mod{n}$, we do the
following. First, we find $q$, the power of 2 with $n^2 \leq q < 2 n^2$.
We will not include $n$, $x$, or $q$ when we write down
the state of our machine, because we never change these values.
In a quantum gate array we need not even keep these values in
memory, as they can be built into the structure of the gate array.
Next, we put the first register
in the uniform superposition of states representing
numbers $a \mod{q}$. This leaves our machine in state
\begin{equation}
\frac{1}{q^{1/2}} \sum_{a=0}^{q-1} \|a\>\|0\>.
\end{equation}
This step is relatively easy, since all it entails is putting each bit
in the first register into the superposition $\frac{1}{\sqrt{2}}(\|0\>
+\|1\>)$.
Next, we compute $x^a\mod{n}$ in the second register as described in~\S3.
Since we keep $a$ in the first register
this can be done reversibly. This leaves our machine in the state
\begin{equation}
\frac{1}{q^{1/2}} \sum_{a=0}^{q-1} \|a\>\|x^a\mod{n}\>.
\end{equation}
We then perform our Fourier transform $A_q$ on the first register,
as described in~\S4,
mapping $\|a\>$ to
\begin{equation}
\frac{1}{q^{1/2}} \sum_{c=0}^{q-1} \exp(2 \pi i ac/q) \|c\>.
\end{equation}
That is, we apply the unitary matrix
with the $(a,c)$ entry equal to $\frac{1}{q^{1/2}} \exp(2 \pi i ac/q)$.
This leaves our machine in state
\begin{equation}
\frac{1}{q} \sum_{a=0}^{q-1} \sum_{c=0}^{q-1}
\exp(2 \pi i ac / q) \|c\>\|x^a\mod{n}\>.
\end{equation}
Finally, we observe the machine. It would be sufficient to observe solely
the value of $\|c\>$ in the first register, but for clarity we will
assume that we observe both $\|c\>$
and $\|x^a\mod{n}\>$.
We now compute the probability that our machine ends in a particular
state $\|c,x^k\mod{n}\>$, where we may assume $0 \leq k < r$.
Summing over all possible ways to reach the state
$\|c,x^k\mod{n}\>$, we find that
this probability is
\begin{equation}
\left|
\frac{1}{q} \sum_{a:\, x^a\equiv x^k}
\exp(2 \pi i ac / q)
\right|^2 .
\end{equation}
where the sum is over all $a$, $0\leq a < q$, such
that $x^a \equiv x^k\mod{n}$.
Because the order of $x$ is~$r$, this sum is over
all $a$ satisfying
$a \equiv k \mod{r}$.
Writing $a=br+k$, we find that the above probability is
\begin{equation}
\left|
\frac{1}{q} \sum_{b=0}^{\left\lfloor (q-k-1)/r \right\rfloor}
\exp(2 \pi i (br+k) c / q)
\right|^2 .
\end{equation}
We can ignore the term of $\exp(2 \pi i kc/q)$, as it can
be factored out of the
sum and has magnitude~1. We can also replace $rc$ with $\lr{rc}{q}$,
where $\lr{rc}{q}$ is the residue which is congruent to $rc\mod{q}$
and is in the range $-q/2 < \lr{rc}{q} \leq q/2$. This leaves us with
the expression
\begin{equation}
\left|
\frac{1}{q} \sum_{b=0}^{\left\lfloor (q-k-1)/r \right\rfloor}
\exp(2 \pi i b \lr{rc}{q} / q)
\right|^2 .
\label{cprobs}
\end{equation}
We will now show that if $\lr{rc}{q}$
is small enough, all the amplitudes in this sum will be
in nearly the same direction (i.e., have close to the same phase),
and thus make the sum large.
Turning the sum into an integral, we obtain
\begin{equation}
\frac{1}{q} \int_{0}^{\left\lfloor \frac{q-k-1}{r} \right\rfloor}
\exp(2 \pi i b \lr{rc}{q} / q) db +
O\left({\textstyle \frac{\lfloor (q-k-1)/r \rfloor}{q}}
\left( \exp(2\pi i \lr{rc}{q} /q) -1 \right) \right).
\end{equation}
If $|\lr{rc}{q}| \leq r/2$, the error term in the above expression is easily
seen to be bounded by $O(1/q)$. We now show that if $|\lr{rc}{q}| \leq r/2$,
the above integral is large,
so the probability of obtaining a state $\|c,x^k\mod{n}\>$ is large.
Note that this condition depends only on $c$ and is independent of~$k$.
Substituting $u=rb/q$ in the above integral,
we get
\begin{equation}
\frac{1}{r} \int_{0}^{\frac{r}{q}\left\lfloor\frac{q-k-1}{r}\right\rfloor}
\exp\left( 2\pi i {\textstyle\frac{\lr{rc}{q}}{r}}u \right) du.
\end{equation}
Since $k < r$, approximating the upper limit of integration by~1 results
in only a $O(1/q)$
error in the above expression. If we do this, we obtain the integral
\begin{equation}
\frac{1}{r} \int_{0}^{1}
\exp\left(2\pi i {\textstyle\frac{\lr{rc}{q}}{r}}u \right) du.
\label{factorint3}
\end{equation}
Letting $\lr{rc}{q}/r$ vary between $-\frac{1}{2}$ and $\frac{1}{2}$,
the absolute magnitude of the integral (\ref{factorint3}) is easily seen
to be minimized when $\lr{rc}{q}/{r} = \pm \frac{1}{2}$, in which case
the absolute value of expression (\ref{factorint3}) is $2/(\pi r)$.
The square of this quantity is a lower bound on
the probability that we see any particular
state $\|c, x^k\mod{n}\>$ with $\lr{rc}{q} \leq r/2$;
this probability is thus asymptotically
bounded below by $4/(\pi^2 r^2)$, and so is at
least $1/3r^2$ for sufficiently large~$n$.
\begin{figure}
\begin{center}
\setlength{\unitlength}{0.240900pt}
\ifx\plotpoint\undefined\newsavebox{\plotpoint}\fi
\sbox{\plotpoint}{\rule[-0.175pt]{0.350pt}{0.350pt}}%
\begin{picture}(1500,900)(0,0)
\small
\sbox{\plotpoint}{\rule[-0.175pt]{0.350pt}{0.350pt}}%
\put(242,158){\makebox(0,0)[r]{0.00}}
\put(244,158){\rule[-0.175pt]{4.818pt}{0.350pt}}
\put(242,263){\makebox(0,0)[r]{0.02}}
\put(244,263){\rule[-0.175pt]{4.818pt}{0.350pt}}
\put(242,368){\makebox(0,0)[r]{0.04}}
\put(244,368){\rule[-0.175pt]{4.818pt}{0.350pt}}
\put(242,473){\makebox(0,0)[r]{0.06}}
\put(244,473){\rule[-0.175pt]{4.818pt}{0.350pt}}
\put(242,577){\makebox(0,0)[r]{0.08}}
\put(244,577){\rule[-0.175pt]{4.818pt}{0.350pt}}
\put(242,682){\makebox(0,0)[r]{0.10}}
\put(244,682){\rule[-0.175pt]{4.818pt}{0.350pt}}
\put(242,787){\makebox(0,0)[r]{0.12}}
\put(244,787){\rule[-0.175pt]{4.818pt}{0.350pt}}
\put(329,113){\makebox(0,0){0}}
\put(329,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(459,113){\makebox(0,0){32}}
\put(459,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(590,113){\makebox(0,0){64}}
\put(590,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(720,113){\makebox(0,0){96}}
\put(720,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(850,113){\makebox(0,0){128}}
\put(850,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(980,113){\makebox(0,0){160}}
\put(980,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(1110,113){\makebox(0,0){192}}
\put(1110,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(1241,113){\makebox(0,0){224}}
\put(1241,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(1371,113){\makebox(0,0){256}}
\put(1371,138){\rule[-0.175pt]{0.350pt}{4.818pt}}
\put(264,158){\rule[-0.175pt]{282.335pt}{0.350pt}}
\put(1436,158){\rule[-0.175pt]{0.350pt}{151.526pt}}
\put(264,787){\rule[-0.175pt]{282.335pt}{0.350pt}}
\put(45,472){\makebox(0,0)[l]{\shortstack{P}}}
\put(850,68){\makebox(0,0){\it c}}
\put(264,158){\rule[-0.175pt]{0.350pt}{151.526pt}}
\put(329,158){\rule[-0.175pt]{0.350pt}{126.232pt}}
\put(333,158){\usebox{\plotpoint}}
\put(337,158){\usebox{\plotpoint}}
\put(341,158){\usebox{\plotpoint}}
\put(345,158){\usebox{\plotpoint}}
\put(349,158){\usebox{\plotpoint}}
\put(354,158){\usebox{\plotpoint}}
\put(358,158){\usebox{\plotpoint}}
\put(362,158){\usebox{\plotpoint}}
\put(366,158){\usebox{\plotpoint}}
\put(370,158){\usebox{\plotpoint}}
\put(374,158){\usebox{\plotpoint}}
\put(378,158){\usebox{\plotpoint}}
\put(382,158){\usebox{\plotpoint}}
\put(386,158){\usebox{\plotpoint}}
\put(390,158){\usebox{\plotpoint}}
\put(394,158){\usebox{\plotpoint}}
\put(398,158){\usebox{\plotpoint}}
\put(402,158){\usebox{\plotpoint}}
\put(406,158){\usebox{\plotpoint}}
\put(411,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(415,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(419,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(423,158){\rule[-0.175pt]{0.350pt}{1.686pt}}
\put(427,158){\rule[-0.175pt]{0.350pt}{4.577pt}}
\put(431,158){\rule[-0.175pt]{0.350pt}{32.281pt}}
\put(435,158){\rule[-0.175pt]{0.350pt}{72.270pt}}
\put(439,158){\rule[-0.175pt]{0.350pt}{6.022pt}}
\put(443,158){\rule[-0.175pt]{0.350pt}{1.927pt}}
\put(447,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(451,158){\rule[-0.175pt]{0.350pt}{0.723pt}}
\put(455,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(459,158){\usebox{\plotpoint}}
\put(463,158){\usebox{\plotpoint}}
\put(467,158){\usebox{\plotpoint}}
\put(472,158){\usebox{\plotpoint}}
\put(476,158){\usebox{\plotpoint}}
\put(480,158){\usebox{\plotpoint}}
\put(484,158){\usebox{\plotpoint}}
\put(488,158){\usebox{\plotpoint}}
\put(492,158){\usebox{\plotpoint}}
\put(496,158){\usebox{\plotpoint}}
\put(500,158){\usebox{\plotpoint}}
\put(504,158){\usebox{\plotpoint}}
\put(508,158){\usebox{\plotpoint}}
\put(512,158){\usebox{\plotpoint}}
\put(516,158){\usebox{\plotpoint}}
\put(520,158){\usebox{\plotpoint}}
\put(524,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(529,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(533,158){\rule[-0.175pt]{0.350pt}{3.132pt}}
\put(537,158){\rule[-0.175pt]{0.350pt}{110.573pt}}
\put(541,158){\rule[-0.175pt]{0.350pt}{6.986pt}}
\put(545,158){\rule[-0.175pt]{0.350pt}{1.445pt}}
\put(549,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(553,158){\usebox{\plotpoint}}
\put(557,158){\usebox{\plotpoint}}
\put(561,158){\usebox{\plotpoint}}
\put(565,158){\usebox{\plotpoint}}
\put(569,158){\usebox{\plotpoint}}
\put(573,158){\usebox{\plotpoint}}
\put(577,158){\usebox{\plotpoint}}
\put(581,158){\usebox{\plotpoint}}
\put(585,158){\usebox{\plotpoint}}
\put(590,158){\usebox{\plotpoint}}
\put(594,158){\usebox{\plotpoint}}
\put(598,158){\usebox{\plotpoint}}
\put(602,158){\usebox{\plotpoint}}
\put(606,158){\usebox{\plotpoint}}
\put(610,158){\usebox{\plotpoint}}
\put(614,158){\usebox{\plotpoint}}
\put(618,158){\usebox{\plotpoint}}
\put(622,158){\usebox{\plotpoint}}
\put(626,158){\usebox{\plotpoint}}
\put(630,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(634,158){\rule[-0.175pt]{0.350pt}{1.445pt}}
\put(638,158){\rule[-0.175pt]{0.350pt}{6.986pt}}
\put(642,158){\rule[-0.175pt]{0.350pt}{110.573pt}}
\put(647,158){\rule[-0.175pt]{0.350pt}{3.132pt}}
\put(651,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(655,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(659,158){\usebox{\plotpoint}}
\put(663,158){\usebox{\plotpoint}}
\put(667,158){\usebox{\plotpoint}}
\put(671,158){\usebox{\plotpoint}}
\put(675,158){\usebox{\plotpoint}}
\put(679,158){\usebox{\plotpoint}}
\put(683,158){\usebox{\plotpoint}}
\put(687,158){\usebox{\plotpoint}}
\put(691,158){\usebox{\plotpoint}}
\put(695,158){\usebox{\plotpoint}}
\put(699,158){\usebox{\plotpoint}}
\put(704,158){\usebox{\plotpoint}}
\put(708,158){\usebox{\plotpoint}}
\put(712,158){\usebox{\plotpoint}}
\put(716,158){\usebox{\plotpoint}}
\put(720,158){\usebox{\plotpoint}}
\put(724,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(728,158){\rule[-0.175pt]{0.350pt}{0.723pt}}
\put(732,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(736,158){\rule[-0.175pt]{0.350pt}{1.927pt}}
\put(740,158){\rule[-0.175pt]{0.350pt}{6.022pt}}
\put(744,158){\rule[-0.175pt]{0.350pt}{72.270pt}}
\put(748,158){\rule[-0.175pt]{0.350pt}{32.281pt}}
\put(752,158){\rule[-0.175pt]{0.350pt}{4.577pt}}
\put(756,158){\rule[-0.175pt]{0.350pt}{1.686pt}}
\put(760,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(765,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(769,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(773,158){\usebox{\plotpoint}}
\put(777,158){\usebox{\plotpoint}}
\put(781,158){\usebox{\plotpoint}}
\put(785,158){\usebox{\plotpoint}}
\put(789,158){\usebox{\plotpoint}}
\put(793,158){\usebox{\plotpoint}}
\put(797,158){\usebox{\plotpoint}}
\put(801,158){\usebox{\plotpoint}}
\put(805,158){\usebox{\plotpoint}}
\put(809,158){\usebox{\plotpoint}}
\put(813,158){\usebox{\plotpoint}}
\put(817,158){\usebox{\plotpoint}}
\put(822,158){\usebox{\plotpoint}}
\put(826,158){\usebox{\plotpoint}}
\put(830,158){\usebox{\plotpoint}}
\put(834,158){\usebox{\plotpoint}}
\put(838,158){\usebox{\plotpoint}}
\put(842,158){\usebox{\plotpoint}}
\put(846,158){\usebox{\plotpoint}}
\put(850,158){\rule[-0.175pt]{0.350pt}{126.232pt}}
\put(854,158){\usebox{\plotpoint}}
\put(858,158){\usebox{\plotpoint}}
\put(862,158){\usebox{\plotpoint}}
\put(866,158){\usebox{\plotpoint}}
\put(870,158){\usebox{\plotpoint}}
\put(874,158){\usebox{\plotpoint}}
\put(878,158){\usebox{\plotpoint}}
\put(883,158){\usebox{\plotpoint}}
\put(887,158){\usebox{\plotpoint}}
\put(891,158){\usebox{\plotpoint}}
\put(895,158){\usebox{\plotpoint}}
\put(899,158){\usebox{\plotpoint}}
\put(903,158){\usebox{\plotpoint}}
\put(907,158){\usebox{\plotpoint}}
\put(911,158){\usebox{\plotpoint}}
\put(915,158){\usebox{\plotpoint}}
\put(919,158){\usebox{\plotpoint}}
\put(923,158){\usebox{\plotpoint}}
\put(927,158){\usebox{\plotpoint}}
\put(931,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(935,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(940,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(944,158){\rule[-0.175pt]{0.350pt}{1.686pt}}
\put(948,158){\rule[-0.175pt]{0.350pt}{4.577pt}}
\put(952,158){\rule[-0.175pt]{0.350pt}{32.281pt}}
\put(956,158){\rule[-0.175pt]{0.350pt}{72.270pt}}
\put(960,158){\rule[-0.175pt]{0.350pt}{6.022pt}}
\put(964,158){\rule[-0.175pt]{0.350pt}{1.927pt}}
\put(968,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(972,158){\rule[-0.175pt]{0.350pt}{0.723pt}}
\put(976,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(980,158){\usebox{\plotpoint}}
\put(984,158){\usebox{\plotpoint}}
\put(988,158){\usebox{\plotpoint}}
\put(992,158){\usebox{\plotpoint}}
\put(997,158){\usebox{\plotpoint}}
\put(1001,158){\usebox{\plotpoint}}
\put(1005,158){\usebox{\plotpoint}}
\put(1009,158){\usebox{\plotpoint}}
\put(1013,158){\usebox{\plotpoint}}
\put(1017,158){\usebox{\plotpoint}}
\put(1021,158){\usebox{\plotpoint}}
\put(1025,158){\usebox{\plotpoint}}
\put(1029,158){\usebox{\plotpoint}}
\put(1033,158){\usebox{\plotpoint}}
\put(1037,158){\usebox{\plotpoint}}
\put(1041,158){\usebox{\plotpoint}}
\put(1045,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(1049,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(1053,158){\rule[-0.175pt]{0.350pt}{3.132pt}}
\put(1058,158){\rule[-0.175pt]{0.350pt}{110.573pt}}
\put(1062,158){\rule[-0.175pt]{0.350pt}{6.986pt}}
\put(1066,158){\rule[-0.175pt]{0.350pt}{1.445pt}}
\put(1070,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(1074,158){\usebox{\plotpoint}}
\put(1078,158){\usebox{\plotpoint}}
\put(1082,158){\usebox{\plotpoint}}
\put(1086,158){\usebox{\plotpoint}}
\put(1090,158){\usebox{\plotpoint}}
\put(1094,158){\usebox{\plotpoint}}
\put(1098,158){\usebox{\plotpoint}}
\put(1102,158){\usebox{\plotpoint}}
\put(1106,158){\usebox{\plotpoint}}
\put(1110,158){\usebox{\plotpoint}}
\put(1115,158){\usebox{\plotpoint}}
\put(1119,158){\usebox{\plotpoint}}
\put(1123,158){\usebox{\plotpoint}}
\put(1127,158){\usebox{\plotpoint}}
\put(1131,158){\usebox{\plotpoint}}
\put(1135,158){\usebox{\plotpoint}}
\put(1139,158){\usebox{\plotpoint}}
\put(1143,158){\usebox{\plotpoint}}
\put(1147,158){\usebox{\plotpoint}}
\put(1151,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(1155,158){\rule[-0.175pt]{0.350pt}{1.445pt}}
\put(1159,158){\rule[-0.175pt]{0.350pt}{6.986pt}}
\put(1163,158){\rule[-0.175pt]{0.350pt}{110.573pt}}
\put(1167,158){\rule[-0.175pt]{0.350pt}{3.132pt}}
\put(1171,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(1176,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(1180,158){\usebox{\plotpoint}}
\put(1184,158){\usebox{\plotpoint}}
\put(1188,158){\usebox{\plotpoint}}
\put(1192,158){\usebox{\plotpoint}}
\put(1196,158){\usebox{\plotpoint}}
\put(1200,158){\usebox{\plotpoint}}
\put(1204,158){\usebox{\plotpoint}}
\put(1208,158){\usebox{\plotpoint}}
\put(1212,158){\usebox{\plotpoint}}
\put(1216,158){\usebox{\plotpoint}}
\put(1220,158){\usebox{\plotpoint}}
\put(1224,158){\usebox{\plotpoint}}
\put(1228,158){\usebox{\plotpoint}}
\put(1233,158){\usebox{\plotpoint}}
\put(1237,158){\usebox{\plotpoint}}
\put(1241,158){\usebox{\plotpoint}}
\put(1245,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(1249,158){\rule[-0.175pt]{0.350pt}{0.723pt}}
\put(1253,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(1257,158){\rule[-0.175pt]{0.350pt}{1.927pt}}
\put(1261,158){\rule[-0.175pt]{0.350pt}{6.022pt}}
\put(1265,158){\rule[-0.175pt]{0.350pt}{72.270pt}}
\put(1269,158){\rule[-0.175pt]{0.350pt}{32.281pt}}
\put(1273,158){\rule[-0.175pt]{0.350pt}{4.577pt}}
\put(1277,158){\rule[-0.175pt]{0.350pt}{1.686pt}}
\put(1281,158){\rule[-0.175pt]{0.350pt}{0.964pt}}
\put(1285,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(1290,158){\rule[-0.175pt]{0.350pt}{0.482pt}}
\put(1294,158){\usebox{\plotpoint}}
\put(1298,158){\usebox{\plotpoint}}
\put(1302,158){\usebox{\plotpoint}}
\put(1306,158){\usebox{\plotpoint}}
\put(1310,158){\usebox{\plotpoint}}
\put(1314,158){\usebox{\plotpoint}}
\put(1318,158){\usebox{\plotpoint}}
\put(1322,158){\usebox{\plotpoint}}
\put(1326,158){\usebox{\plotpoint}}
\put(1330,158){\usebox{\plotpoint}}
\put(1334,158){\usebox{\plotpoint}}
\put(1338,158){\usebox{\plotpoint}}
\put(1342,158){\usebox{\plotpoint}}
\put(1346,158){\usebox{\plotpoint}}
\put(1351,158){\usebox{\plotpoint}}
\put(1355,158){\usebox{\plotpoint}}
\put(1359,158){\usebox{\plotpoint}}
\put(1363,158){\usebox{\plotpoint}}
\put(1367,158){\usebox{\plotpoint}}
\end{picture}
\end{center}
\caption{The probability $\rm P$ of observing values of $c$ between
$0$ and $255$, given
$q=256$ and $r=10$.}
\label{probabilityplot}
\end{figure}
The probability of seeing a given
state $\|c,x^k\mod{n}\>$ will thus be at least $1/3r^2$ if
\begin{equation}
\frac{-r}{2} \leq \lr{rc}{q} \leq \frac{r}{2},
\end{equation}
i.e., if there is a $d$ such that
\begin{equation} \frac{-r}{2} \leq rc-dq \leq \frac{r}{2}. \end{equation}
Dividing by $rq$ and rearranging the terms gives
\begin{equation}
\left| \frac{c}{q} - \frac{d}{r} \right| \leq \frac{1}{2q}.
\end{equation}
We know $c$ and~$q$.
Because $q > n^2$, there is at most one fraction $d/r$ with $r<n$
that satisfies the above inequality.
Thus, we can obtain the fraction $d/r$ in lowest terms by rounding
$c/q$ to the nearest fraction having a denominator smaller than~$n$.
This fraction can be found in polynomial time by using a continued fraction
expansion of $c/q$, which finds all the best approximations of
$c/q$ by fractions [Hardy and Wright 1979, Chapter X, Knuth 1981].
The exact probabilities as
given by equation (\ref{cprobs}) for an example case with $r=10$ and
$q=256$ are plotted in
Figure~\ref{probabilityplot}. The value $r=10$ could occur when factoring
33 if $x$ were chosen to be~5, for example. Here $q$ is taken smaller than
$33^2$ so as to make the values of $c$ in the plot distinguishable;
this does not change the functional structure of ${\rm P}(c)$.
Note that with high probability the
observed value of $c$ is near an integral multiple of $q/r = 256/10$.
If we have the fraction $d/r$ in lowest terms, and if $d$ happens to be
relatively prime to~$r$, this will give us~$r$.
We will now count the number of states $\|c,x^k\mod{n}\>$ which
enable us to compute $r$ in this way. There are $\phi(r)$ possible values
of $d$ relatively prime to~$r$, where $\phi$ is Euler's totient function
[Knuth 1981, Hardy and Wright 1979, \S5.5].
Each of these fractions $d/r$ is close to one fraction $c/q$ with
$|c/q-d/r| \leq 1/2q$. There are also $r$ possible values for~$x^k$,
since $r$ is the order of~$x$. Thus, there are $r \phi(r)$ states
$\|c,x^k\mod{n}\>$ which would enable us to obtain~$r$. Since each
of these states occurs with probability at least $1/3r^2$, we obtain
$r$ with probability at least $\phi(r)/3r$. Using the theorem that
$\phi(r)/r > \delta/\log \log r$ for some constant $\delta$
[Hardy and Wright 1979, Theorem 328], this shows that we find $r$ at least a
$\delta/\log\log r$ fraction of the time, so by repeating this experiment
only $O(\log\log r)$ times, we are assured of a high probability of
success.
In practice, assuming that quantum computation is more expensive than
classical computation, it would be worthwhile to alter the above algorithm
so as to perform less quantum computation and
more postprocessing. First, if the observed state is $\|c\>$,
it would be wise to also try numbers close to $c$ such as $c\pm 1$, $c\pm 2$,
$\ldots$, since these also have a reasonable chance of being close to a
fraction $qd/r$. Second, if $c/q \approx d/r$, and $d$ and $r$ have a
common factor, it is likely to be small. Thus,
if the observed value of $c/q$ is rounded off to
$d'/r'$ in lowest terms, for a
candidate $r$ one should consider not only~$r'$ but also its small multiples
$2r'$, $3r'$, \ldots, to see if these are the actual order of~$x$.
Although the first technique will only reduce the expected number of trials
required to find $r$ by a constant factor,
the second technique will reduce the expected number of trials
for the hardest $n$ from $O(\log \log n)$ to $O(1)$
if the first $(\log n)^{1+\epsilon}$ multiples of~$r'$ are considered
\cite{Odly}. A third technique is, if
two candidate $r$'s have been found, say
$r_1$ and~$r_2$, to test the least common multiple of $r_1$ and $r_2$
as a candidate~$r$. This third technique is also able to reduce the
expected number of trials to a constant \cite{Knil}, and will also work in
some cases where the first two techniques fail.
Note that in this algorithm for determining the order of an
element, we did not use many of the
properties of multiplication $\bmod{n}$. In fact, if we have a permutation
$f$ mapping the set $\{0,1,2,\ldots,n-1\}$ into itself such
that its $k$th iterate, $f^{(k)}(a)$, is computable in time polynomial
in $\log n$ and $\log k$,
the same algorithm will be able to find the order of an element $a$
under~$f$, i.e., the minimum $r$ such that $f^{(r)}(a)=a$.
\section{Discrete logarithms}
For every prime~$p$, the multiplicative group $\bmod{p}$ is cyclic,
that is, there are generators $g$ such that $1$, $g$, $g^2$, \ldots,
$g^{p-2}$
comprise all the non-zero residues $\bmod{p}$ [Hardy and Wright 1979,
Theorem 111, Knuth 1981]. Suppose we are given a prime $p$ and such a
generator~$g$. The {\em discrete logarithm} of a number $x$ with
respect to $p$ and $g$ is the integer $r$ with $0 \leq r < p-1$ such that
$g^r \equiv x \mod{p}$. The fastest algorithm known for finding discrete
logarithms modulo arbitrary primes $p$ is Gordon's [1993] adaptation of
the number field sieve, which runs in time
$\exp(O(\log p)^{1/3} (\log \log p)^{2/3}))$. We show how to find discrete
logarithms on a quantum computer with two
modular exponentiations and two quantum Fourier transforms.
This algorithm will use three quantum registers. We first
find $q$ a power of 2 such that $q$ is close to~$p$, i.e., with
$p < q < 2p$.
Next, we put the first two registers in
our quantum computer in the uniform superposition of
all $\|a\>$ and $\|b\>$
$\bmod{p-1}$, and compute $g^ax^{-b}\mod{p}$ in the third register. This
leaves our machine in the state
\begin{equation}
\frac{1}{p-1}\sum_{a=0}^{p-2}\sum_{b=0}^{p-2} \|a,b,g^a x^{-b}\mod{p}\>.
\end{equation}
As before, we use the Fourier transform $A_q$ to
send $\|a\>\rightarrow \|c\>$ and $\|b\>\rightarrow \|d\>$
with probability amplitude $\frac{1}{q}\exp(2\pi i (ac+bd)/q)$.
This is, we take the state $\|a,b\>$ to the state
\begin{equation}
\frac{1}{q} \sum_{c=0}^{q-1} \sum_{d=0}^{q-1}
\textstyle \exp\big(\frac{2 \pi i}{q} (ac+bd)\big)
\|c,d\>.
\end{equation}
This leaves our quantum computer in the state
\begin{equation}
{\frac{1}{(p-1)q}}
\sum_{a,b=0\ }^{p-2} \sum_{\ c,d=0}^{q-1}
\textstyle \exp\big(\frac{2 \pi i}{q} (ac+bd)\big)
\|c,d,g^a x^{-b}\!\mod{p}\>.
\end{equation}
Finally, we observe the state of the quantum computer.
The probability of observing a state $\|c,d,y\>$ with
$y \equiv g^k \mod{p}$ is
\begin{equation}
\left| \; \frac{1}{(p-1)q}\sum_{{a,b}\atop{a-rb \equiv k}}
\exp\left({\textstyle \frac{2\pi i}{q} } (ac+bd)\right) \;
\right|^2
\label{probobserve}
\end{equation}
where the sum is over all
$(a,b)$ such that $a-rb \equiv k \mod{p-1}$.
Note that we now have two moduli to deal with, $p-1$ and~$q$. While this
makes keeping track of things more confusing, it does not pose serious
problems.
We now use the relation
\begin{equation}
a = br + k - (p-1) \left\lfloor{\textstyle\frac{br+k}{p-1}}\right\rfloor
\label{arelation}
\end{equation}
and substitute (\ref{arelation}) in the expression (\ref{probobserve})
to obtain the amplitude
on $\|c,d,g^k\mod{p}\>$, which is
\begin{equation}
{\frac{1}{(p-1)q}}
\sum_{b=0}^{p-2}
\exp\Big(\textstyle\frac{2\pi i}{q} \big(brc+kc+bd-c(p-1)\left\lfloor
{\textstyle\frac{br+k}{p-1}}\right\rfloor\big)\Big).
\label{amplitudeexpress}
\end{equation}
The absolute value of the square of this amplitude is the probability
of observing the state $\|c,d,g^k\mod{p}\>$.
We will now analyze the expression
(\ref{amplitudeexpress}). First, a factor of
$\exp(2\pi i k c/q)$ can be taken out of all the terms and ignored,
because it does not change the probability. Next, we split the
exponent into two parts and factor out $b$ to obtain
\begin{equation}
\frac{1}{(p-1)q}\sum_{b=0}^{p-2}
\textstyle
\exp\left(\frac{2\pi i}{q}
bT\right)
\exp\left(\frac{2\pi i}{q}
V\right),
\label{bigeq}
\end{equation}
where
\begin{equation}
T = \textstyle {rc+d-\frac{r}{p-1}\lr{c(p-1)}{q}},\!\phantom{pen}
\end{equation}
and
\begin{equation}
V = \textstyle{\left(\frac{br}{p-1}-
\left\lfloor\frac{br+k}{p-1}\right\rfloor\right)\lr{c(p-1)}{q}}.
\end{equation}
Here by $\lr{z}{q}$ we mean the residue of $z \mod q$ with
$-q/2 < \lr{z}{q} \leq q/2$, as in equation (\ref{cprobs}).
We next classify possible outputs (observed states) of the quantum
computer into ``good'' and ``bad.''
We will show that if we get enough ``good'' outputs, then we will likely
be able to
deduce~$r$, and that furthermore, the chance of getting a ``good''
output is constant. The idea is that if
\begin{equation}
\big|\lr{T}{q}\big| =
\big| rc+d-{\textstyle \frac{r}{p-1}}\lr{c(p-1)}{q} -jq \big|
\leq \frac{1}{2},
\label{firstcond}
\end{equation}
where $j$ is the closest integer to $ T/q $,
then as $b$ varies between 0 and $p-2$, the phase of
the first exponential term in equation (\ref{bigeq}) only varies
over at most half of the unit circle. Further, if
\begin{equation}
\left| \lr{c(p-1)}{q} \right| \leq q/12,
\label{secondcond}
\end{equation}
then $|V|$ is always at most $q/12$, so the phase of
the second exponential term in equation (\ref{bigeq}) never is farther than
$\exp(\pi i/ 6)$ from~1. If conditions (\ref{firstcond}) and
(\ref{secondcond}) both hold, we will say that an output is
``good.'' We will
show that if both conditions hold, then the contribution to the
probability from the corresponding term is significant. Furthermore,
both conditions will hold with constant probability, and
a reasonable sample of $c$'s for which condition (\ref{firstcond}) holds
will allow us to deduce~$r$.
We now give a lower bound
on the probability of each good output, i.e., an output that satisfies
conditions (\ref{firstcond}) and (\ref{secondcond}).
We know that as $b$ ranges from 0 to $p-2$, the phase of
$\exp(2\pi i bT/q)$ ranges from
$0$ to $2\pi i W$ where
\begin{equation}
W = \frac{p-2}{q}
{\left(rc+d-\frac{r}{p-1}\lr{c(p-1)}{q} -jq \right)}
\end{equation}
and $j$ is as in equation (\ref{firstcond}).
Thus, the component of the
amplitude of the first exponential in the summand of (\ref{bigeq})
in the direction
\begin{equation}
\exp\left( \pi i W
\right)
\label{direction}
\end{equation}
is at least $\cos(2 \pi \left| W/2 - Wb/(p-2)\right|)$. By condition
(\ref{secondcond}), the phase can vary by at most ${\pi i/6}$ due to
the second exponential $\exp(2\pi i V/q)$.
Applying this variation in the manner that minimizes the component
in the direction (\ref{direction}), we get that the component in this
direction is at least
\begin{equation}
\cos(2 \pi \left| W/2 - Wb/(p-2)\right| + {\textstyle\frac{\pi}{6}}).
\end{equation}
Thus we get that the absolute value of the amplitude (\ref{bigeq}) is at
least
\begin{equation}
\frac{1}{(p-1)q} \sum_{b=0}^{p-2} \cos\left(2 \pi \left| W/2-Wb/(p-2)\right|
+ {\textstyle\frac{\pi}{6}}\right).
\end{equation}
Replacing this sum with an integral, we get that
the absolute value of this amplitude is at least
\begin{equation}
\frac{2}{q} \int_{0}^{1/2} \cos (\textstyle \frac{\pi}{6} + 2\pi |W| u) du
\ + \ O\left(\frac{W}{pq}\right).
\label{logintegral1}
\end{equation}
From condition (\ref{firstcond}), $|W| \leq \frac{1}{2}$, so the error term
is $O(\frac{1}{pq})$.
As $W$ varies between $-\frac{1}{2}$ and $\frac{1}{2}$,
the integral (\ref{logintegral1})
is minimized when $|W|=\frac{1}{2}$.
Thus, the probability of
arriving at a state $\|c,d,y\>$ that satisfies both conditions
(\ref{firstcond}) and (\ref{secondcond}) is at least
\begin{equation}
\left(
\frac{1}{q} \frac{2}{\pi} \int_{\pi/6}^{2\pi/3} \cos u \; du
\right)^2,
\end{equation}
or at least $.054 / q^2 > 1/(20q^2)$.
We will now count the number of pairs $(c,d)$ satisfying conditions
(\ref{firstcond}) and (\ref{secondcond}).
The number of pairs $(c,d)$ such
that (\ref{firstcond}) holds is exactly the number of possible
$c$'s, since
for every $c$ there is exactly one $d$ such that (\ref{firstcond})
holds. Unless $\gcd(p-1,q)$ is large, the number of $c$'s
for which (\ref{secondcond}) holds is approximately $q/6$, and even if
it is large, this number is at least $q/12$.
Thus,
there are at least $q/12$ pairs $(c,d)$ satisfying
both conditions. Multiplying by $p-1$,
which is the number of possible $y$'s, gives approximately
$pq/12$ good states $\|c,d,y\>$. Combining this calculation with the lower
bound $1/(20q^2)$ on the probability of observing
each good state gives us that the
probability of observing some good state is at least $p/(240q)$, or at
least $1/480$ (since $q < 2p$). Note that each good
$c$ has a probability of at least $(p-1)/(20q^2) \geq 1/(40q)$ of being
observed, since there $p-1$ values of $y$ and one value of $d$ with
which $c$ can make a good state $\|c,d,y\>$.
We now want to recover $r$ from a pair $c,d$ such that
\begin{equation}
-\frac{1}{2q} \leq
\frac{d}{q} + r\left(\frac{c(p-1)-\lr{c(p-1)}{q}}{(p-1)q}\right) \leq
\frac{1}{2q} \ \ \ \ \mod{1},
\end{equation}
where this equation was obtained from condition (\ref{firstcond}) by
dividing by~$q$. The first thing to notice is that the multiplier on~$r$
is a fraction with denominator $p-1$, since $q$ evenly divides
$c(p-1) - \lr{c(p-1)}{q}$. Thus, we
need only round $d/q$ off to the nearest multiple of $1/(p-1)$ and divide
$\bmod{p-1}$ by the integer
\begin{equation}
c' = \frac{c(p-1)-\lr{c(p-1)}{q}}{q}
\label{definecprime}
\end{equation}
to find a candidate~$r$. To show that the quantum calculation need
only be repeated a polynomial number of times to find the correct~$r$
requires only a few more details. The problem is that we cannot
divide by a number $c'$ which is not relatively prime to $p-1$.
For the
discrete log algorithm, we do not know that all possible values
of $c'$ are generated with reasonable likelihood; we only know this about
one-twelfth of them. This additional difficulty makes the next step harder
than the corresponding step in the algorithm for factoring. If we knew the
remainder of $r$ modulo all prime powers dividing $p-1$, we could use the
Chinese remainder theorem to recover $r$ in polynomial time. We will only
be able to prove that we can find this remainder
for primes larger than~18, but with a little extra work
we will still be able to recover~$r$.
Recall that each good $(c,d)$ pair is generated with probability
at least $1/(20q^2)$, and that at least a twelfth of the possible $c$'s
are in a good $(c,d)$ pair. From equation (\ref{definecprime}), it follows
that these $c$'s are mapped from $c/q$ to
$c'/(p-1)$ by rounding to the nearest integral multiple of \linebreak[3]
$1/\linebreak[1](p-1)$.
Further, the good $c$'s are exactly those in which ${c}/{q}$ is
close to ${c'}/(p-1)$. Thus, each good $c$ corresponds with exactly
one~$c'$. We would like to show that for any prime power
$p_i^{\alpha_i}$ dividing $p-1$, a random good $c'$ is unlikely to
contain~$p_i$. If we are willing to accept a large constant for
our algorithm, we can just ignore the prime powers under~18; if we know
$r$ modulo all prime powers
over~18, we can try all possible residues for primes under 18 with only a
(large) constant factor increase in running time.
Because at least one twelfth of the $c$'s were in a good
$(c,d)$ pair, at least one twelfth of the $c'$'s are good. Thus,
for a prime power $p_i^{\alpha_i}$, a random good $c'$ is
divisible by $p_i^{\alpha_i}$ with probability at most $12/p_i^{\alpha_i}$.
If we have $t$ good $c'$'s, the probability
of having a prime power over 18 that divides all of them is therefore at most
\begin{equation}
\sum_{18 \,<\, p_i^{\alpha_i} \big| (p-1)}
\left(\frac{12}{p_i^{\alpha_i}}\right)^t,
\end{equation}
where $a|b$ means that $a$ evenly divides~$b$, so
the sum is over all prime powers greater than 18 that divide $p-1$.
This sum (over all integers $> 18$) converges for $t=2$, and goes down
by at least a factor of $2/3$ for each further increase of $t$ by~1; thus for
some constant $t$ it is less than $1/2$.
Recall that each good $c'$
is obtained with probability at least $1/(40q)$ from any experiment. Since
there are $q/12$ good $c'$'s, after $480t$ experiments, we are likely
to obtain a sample of $t$ good $c'$'s chosen equally likely from all
good $c'$'s. Thus, we will be able to find a set of $c'$'s such
that all prime powers $p_i^{\alpha_i} > 20$ dividing $p-1$ are relatively
prime to at least one of these $c'$'s.
To obtain a polynomial time algorithm,
all one need do is try all possible sets of $c'$'s of size~$t$; in practice,
one would use an algorithm to find sets of $c'$'s with large common factors.
This set gives the residue of $r$ for all primes larger than~18.
For each prime $p_i$ less than~18, we have at most 18 possibilities
for the residue modulo $p_i^{\alpha_i}$, where $\alpha_i$ is the exponent
on prime $p_i$ in the prime factorization of $p-1$. We can thus try all
possibilities for residues modulo powers of primes less than~18: for each
possibility we can calculate the corresponding~$r$ using the Chinese
remainder theorem and then check to see whether it is the desired discrete
logarithm.
If one were to actually program this algorithm
there are many ways in which the efficiency
could be increased over the efficiency shown in this paper.
For example, the estimate for the number of good $c'$'s is likely too low,
especially since weaker conditions than (\ref{firstcond}) and
(\ref{secondcond}) should suffice.
This means that the number of times the experiment need be run could be
reduced. It also
seems improbable that the distribution of bad values of $c'$ would have any
relationship to primes under~18; if this is true,
we need not treat small prime powers separately.
This algorithm does not use very many properties of ${\rm Z}_p$, so we can
use the same algorithm to find discrete logarithms over other fields
such as $Z_{p^\alpha}$, as long as the field has a cyclic multiplicative
group. All we need is that we know the order of the
generator, and that we can multiply and take inverses of elements in
polynomial time. The order of the generator could in fact be computed
using the quantum order-finding algorithm given in \S5 of this paper.
Boneh and Lipton [1995] have generalized the algorithm so as to be able to
find discrete logarithms when the group is abelian but not cyclic.
\section{Comments and open problems}
It is currently believed that the most difficult aspect
of building an actual quantum computer will be dealing with the problems of
imprecision and decoherence. It was shown by Bennett et al.
[1994] that the quantum gates need only have precision
$O(1/t)$ in order to have a reasonable probability of completing $t$ steps
of quantum computation; that is, there is a $c$ such that
if the amplitudes in the unitary matrices
representing the quantum gates are all perturbed by at most $c/t$, the
quantum computer will still have a reasonable chance of producing the
desired output. Similarly, the decoherence needs to be only polynomially
small in $t$ in order to have a reasonable probability of completing
$t$ steps of computation successfully. This holds not only for the simple
model of decoherence where each bit has a fixed probability of
decohering at each time step, but also for more complicated models of
decoherence which are derived from fundamental quantum mechanical
considerations \cite{Unru, PaSuEk, ChLaShZu}. However, building
quantum computers with high enough precision and low enough decoherence
to accurately perform long computations may present formidable
difficulties to experimental physicists. In classical computers, error
probabilities can be reduced not only though hardware but also through
software, by the use of redundancy and error-correcting codes. The most
obvious method of using redundancy in quantum computers is ruled out by
the theorem that quantum bits cannot be cloned [Peres 1993, \S\mbox{9-4}],
but this argument does
not rule out more complicated ways of reducing inaccuracy or decoherence
using software. In fact, some progress in the direction of reducing
inaccuracy \cite{BeDeJo} and decoherence \cite{Shor2}
has already been made.
The result of Bennett et al.\ [1995] that quantum bits can
be faithfully transmitted over a noisy quantum channel gives further
hope that quantum computations can similarly be faithfully carried out
using noisy quantum bits and noisy quantum gates.
Discrete logarithms and factoring are not in themselves widely useful
problems. They have only become useful because they have been
found to be crucial for public-key cryptography, and this application
is in turn possible only because they have been presumed to be difficult.
This is also true of the generalizations of Boneh and Lipton [1995] of
these algorithms.
If the only uses of quantum computation remain discrete logarithms and
factoring, it will likely become a
special-purpose technique whose only {\em raison d'\^etre}
is to thwart public key cryptosystems. However, there may be other
hard problems which could be solved asymptotically faster with quantum
computers. In particular, of interesting problems not known
to be NP-complete, the problem of finding a short vector in a lattice
\cite{Adle,AdMc} seems as if it might potentially be
amenable to solution by a quantum computer.
In the history of computer science, however, most important problems
have turned out to be either polynomial-time or NP-complete. Thus
quantum computers will likely not become widely useful unless they can
solve NP-complete problems. Solving NP-complete problems efficiently
is a Holy Grail of theoretical computer science which very few people expect
to be possible on a classical computer. Finding polynomial-time algorithms
for solving these problems on a quantum computer would be a momentous
discovery. There are some weak indications that
quantum computers are not powerful enough to solve NP-complete problems
\cite{BeBeBrVa},
but I do not believe that this potentiality should be ruled out as yet.
\section*{Acknowledgements}
I would like to thank Jeff Lagarias for finding and fixing a critical error
in the first version of the discrete log algorithm. I would also like
to thank him, David Applegate,
Charles Bennett, Gilles Brassard, Andrew Odlyzko, Dan Simon,
Bob Solovay, Umesh Vazirani, and correspondents too numerous
to list, for productive discussions, for corrections to and improvements of
early drafts of this paper, and for pointers to the literature.
{
|
\section{INTRODUCTION}
The liquid alkali metals have been studied extensively in both experimental
and theoretical sides. They can be easily used to test a theoretical
approach as the first step, since they constitute ``simple" metals and
``simple" liquids: furthermore, there exist many reliable experimental
results to be compared. In the standard theory, a liquid metal is treated
as a one-component liquid interacting via a binary effective potential,
which is determined by the pseudopotential formalism; a pseudopotential
is introduced either by first-principles calculations or by adjusting
parameters involved in model potentials to some experimental results.
In this treatment, the ionic structures are determined independently of
the electronic structures in a liquid metal.
It is only recent that a
liquid metal is thought of as an electron-ion mixture and the ionic
structures are determined in a coupled manner with
the electronic structures. One of such approaches is the
Car-Parrinello molecular-dynamics (CP-MD) technique\cite{CPMD},
where a liquid metal is taken as a binary mixture of ions
and electrons. In the CP-MD method, the electron-ion interaction is
described by a pseudopotential to produce pseudo-wavefunctions which
can be accurately represented by small number of plain waves. The CP-MD
method possesses advantage to avoid the difficult task of constructing an
effective ion-ion potential required to perform the molecular-dynamics
simulation, and afford to provide {\it ab initio} calculations of metallic
systems in principle. However, the most serious problem in this approach is
that the number of particles used in the simulations cannot be taken large a
nd
a total of time steps performed in the simulations is limited to a small size
within the present computational resources. In the present work, we
propose another scheme of {\it first principles} molecular-dynamics
simulation based on
the density-functional (DF) method applied to the ion-electron mixture;
this simulation method can be performed on the large number of particles
to last a large size of time steps, since this scheme reduces the electron-i
on
problem to a usual classical MD coupled with a set of integral equations
determining an effective ion-ion interaction: a problem to
determine the ion-configuration structure and the ion-ion
interaction in a self-consistent manner.
Previously, we have proposed a set of integral equations for radial
distribution functions (RDF's) in an electron-ion mixture on the basis
of the DF theory in the quantal hyper-netted (QHNC)
approximation\cite{hyd,QHNC}. In this QHNC formalism, the bare electron-ion
interaction $v_{\rm eI}(r)$ and the ionic charge $Z_{\rm I}$ are
determined self-consistently by regarding a liquid metal as a mixture
of nuclei and electrons\cite{NEmodel}. Already, we have
applied this approach to
liquid metallic hydrogen\cite{hyd}, lithium\cite{QHLi},
sodium\cite{QHNa}, potassium\cite{QHK} and aluminum\cite{QHAl} obtaining
ion-ion structure factors in excellent agreements with experiments.
The QHNC equations are derived from exact expressions for the electron-ion
and ion-ion RDF's in an electron-ion mixture: these exact expressions
are only formal results derived from the DF theory.
A molecular-dynamics scheme to treat
an ion-electron mixture can be set up on the basis of these exact
relations, which states that the electron-ion mixture can be regarded
as a quasi-one component liquid interacting via a ${\it pairwise}$
interaction in the description of the ion-ion RDF\cite{QHNC}
(hereafter, referred to as the QHNC-MD method).
Since the QHNC formalism is derived on the electron-ion model
where the bound-electrons forming an ion in a liquid metal is
assumed to be clearly distinguished from the conduction electrons
and the overlap of the core electrons is negligible, the QHNC-MD
method is only applicable to simple liquid metals: its application
to liquid alkali metals is taken as an ideal test of the QHNC-MD method.
Thus, the QHNC-MD method has been shown to
yield structure factors of liquid alkali metals
in excellent agreement with experiments as the result of a
first-principles calculation in the present work.
In Sec.~\ref{sec:hnc}, we sketch the QHNC formulation: exact expressions for
RDF's in an electron-ion mixture are obtained from the DF
method\cite{BR}, and the nucleus-electron model is shown to
provide a bare electron-ion interaction, which should be
determined self-consistently. The procedure to perform the MD simulation
based on the QHNC theory is shown in Sec.~\ref{sec:md}: in the QHNC
formulation the effective ion-ion interaction used in the MD simulation
depends on the ionic structure specified by the ion-ion RDF.
Therefore, in the application
of this MD scheme it is important to extrapolate the MD RDF
beyond the truncation radius of the simulation correctly so
as to be used in the determination of
effective ion-ion interaction; this is exemplified by the method
described in Sec.~\ref{sec:md}. Numerical procedure of the QHNC-MD
method and the results of its application to alkali liquid metals
are described in Sec.~\ref{sec:appl}.
The last section is devoted to a discussion, where
the advantage/disadvantage
of the QHNC-MD method against the CP-MD method is argued also.
\section{QUANTAL HYPER-NETTED CHAIN THEORY}
\label{sec:hnc}
A simple liquid metal can be thought of as a binary mixture of ions
with a definite ionic charge $Z_{\rm I}$ and the conduction electrons;
the interactions $v_{ij}(r)$ between particles [$i,j=$I or e]
are taken as pair-wise. The ions constitute a classical fluid,
while the conduction electrons form a quantum fluid.
Let us refer to this mixture as the ion-electron model for a liquid metal.
Since the ions are regarded as classical particles in the electron-ion
model, the ion-ion and electron-ion RDF's become identical with the
ion- and electron-density distributions around a fixed ion in the mixture,
respectively\cite{QHNC}. Because a fixed ion causes external potentials
acting on ions and electrons in the homogeneous mixture,
the DF theory can give the exact expressions
for the ion- and electron-density distributions, $n_{\rm I}(r|{\,\rm I})$
and $n_{\rm e}(r|{\,\rm I})$,
in terms of those of noninteracting systems $n^0_i(r)$ under effective
external potentials $U_i^{\rm eff}(r)$ [$i$=I, e]
\begin{equation}
U_i^{\rm eff}(r)=v_{i{\rm I}}(r)+{\delta {\cal F}_{\rm int}\over\delta
n_i(r|\,{\rm I})}-\mu_i^{\rm int} \label{eq:gvef1}\\
\end{equation}
with the use of ${\cal F}_{\rm int}$ and $\mu_i^{\rm int}$, the
interaction part of the intrinsic free-energy and the chemical
potential, respectively\cite{QDCF}.
As a result, the DF theory provides exact, but formal, expressions for
the ion-ion RDF $g_{\scriptscriptstyle\rm II}(r)$ and electron-ion
RDF $g_{\scriptscriptstyle\rm eI}(r)$ as follows:
\begin{eqnarray}
n_0^{\rm I}g_{\scriptscriptstyle\rm II}(r) & = & n_{\rm I}(r|\,{\rm I})=
n_{\rm I}^0(r|U_{\rm I}^{\rm eff})
\equiv n_0^{\rm I}\exp[-\beta U_{\rm I}^{\rm eff}(r)]\;,\label{eq:gII}\\
n_0^{\rm e}g_{\scriptscriptstyle\rm eI}(r) & = & n_{\rm e}(r|\,{\rm I})=
n_{\rm e}^0(r|U_{\rm e}^{\rm eff})
\equiv\sum_i{|\psi_i(r)|^2\over
\exp[\beta(\varepsilon_i-\mu_0^{\rm e})]+1}\;,\label{eq:geI}
\end{eqnarray}
where $\mu_0^{\rm e}$ denotes the chemical potential of a
non-interacting electron gas, $n_0^{\rm I}$ ($n_0^{\rm e}$) is the
number density of ions (electrons), and $\beta=(k_{\rm B}T)^{-1}$
the inverse temperature. The electron-density distribution $n_{\rm
e}^0(r|U)$ is determined by solving the wave equation for an
electron under the external potential $U(r)$
\begin{equation}
\left[ -(\hbar^2/2m)\nabla^2+U(r) \right]\psi_i(r)
=\varepsilon_i\psi_i(r)\;.
\end{equation}
In the similar way to the case of classical binary mixtures, the effective
external potentials $U_i^{\rm eff}(r)$ given by Eq.~(\ref{eq:gvef1})
are written as
\begin{eqnarray}
U_i^{\rm eff}(r) & = &
v_{i\rm I}(r)-{\it\Gamma}_{i\rm I}(r)/\beta
-B_{i\rm I}(r)/\beta\;,\label{eq:Ueff}\\
{\it\Gamma}_{i\rm I}(r) & \equiv &
\sum_l \int C_{il}(|{\bf r}-{\bf r}'|)n_0^l
[g_{{\scriptscriptstyle l}{\rm I}}(r)-1]d{\bf r}'\;,
\end{eqnarray}
in terms of the direct correlation
functions (DCF's) $C_{ij}(r)$ and the bridge functions $B_{i\rm
I}(r)$. Here, the DCF's $C_{ij}(r)$ in the
ion-electron mixture are defined within the framework of the DF theory
by
\begin{equation}
C_{ij}(|{\bf r}-{\bf r'}|) \equiv -\beta { \delta^2 {\cal F}_{\rm
int}[n_I,n_e] \over \delta n_{i}({\bf r})\delta n_{j}({\bf r'})
}\biggl |_0\;, \label{e:dcf}
\end{equation}
where the suffix 0 denotes the functional derivative at the uniform
densities \cite{QDCF}.
Actually the explicit expression for the DCF's are given by the
Fourier transform in the matrix form
\begin{equation}
\sqrt{{\cal N}}C(k)\sqrt{{\cal N}} = (\widetilde{\chi}_Q^0)^{-1} -
(\widetilde{\chi}_Q)^{-1} \label{e:dcfq}
\end{equation}
in terms of the density response functions, $\widetilde{\chi}_Q\equiv
\parallel \chi_{_{ij}}(k)\parallel$ and $\widetilde{\chi}_Q^0\equiv
\parallel
\chi^{0i}(Q)\delta_{ij}\parallel$, of the interacting and
noninteracting systems, respectively, with ${\cal N}\equiv \parallel
n_0^{i}\delta_{ij}\parallel$\cite{QDCF}.
Note here that the density-density response functions $\chi_{iI}(Q)$
concerning ion reduce to the structure factors $S_{iI}(Q)$ and
$\chi^{0I}(Q)=1$, since the ions behave as classical particles\cite{QHNC}.
>From this definition of the DCF's the Ornstein-Zernike relations are
derived for the ion-electron mixture:
\begin{eqnarray}
g_{\scriptscriptstyle\rm II}(r) & = &
C_{\rm II}(r)+{\it\Gamma}_{\rm II}(r)\;,\label{eq:gIIOZ}\\
g_{\scriptscriptstyle\rm eI}(r) & = &
\hat BC_{\rm eI}(r)+\hat B{\it\Gamma}_{\rm eI}(r)\;.\label{eq:geIOZ}
\end{eqnarray}
Here, $\hat B$ denotes an operator defined by
\begin{equation}
{\cal F}_Q[\hat B^\gamma f(r)]
\equiv (\chi_Q^0)^\gamma{\cal F}_Q[f(r)]
=(\chi_Q^0)^\gamma\int\exp[i{\bf Q}\cdot{\bf r}]f(r)d{\bf r}\;,
\end{equation}
for an arbitrary real number $\gamma$, and represents a
quantum-effect of the electron through the density response
function $\chi_Q^0$ of the non-interacting electron gas.
A set of integral equations (\ref{eq:gII}) and (\ref{eq:geI}) are
exact but formal expressions, as well as all other equations
in the above. However, the ionic charge $Z_{\rm I}$ and
the electron-ion interaction $v_{\rm eI}(r)$ must be given
beforehand, when we apply these formula to a
liquid metal as an ion-electron mixture.
In order to determine these quantities from first principle, a liquid metal
must be treated more fundamentally as a mixture of nuclei and electrons
(the nucleus-electron model), where all interactions between particles are
known as pure Coulombic. In this model, input data
in dealing with a liquid metal
is only the atomic number $Z_{\rm A}$ to specify the material.
For this purpose, let us consider a liquid metal as a mixture of
$N_{\rm I}$ nuclei and $Z_{\rm A}N_{\rm I}$ electrons,
and solve the problem to determine
the electron-density distribution around a nucleus fixed at the origin in
this mixture. Since a fixed nucleus causes an external potential
$U(r)=-Z_{\rm A}e^2/r$ for this mixture to produce an inhomogeneous system,
the DF theory can be applied to this problem. It should be noticed that
DF theory contains some arbitrariness in the choice of a reference
system to describe the system\cite{QDCF}. We can get a simple
description of the nucleus-electron mixture if the reference system is
chosen to be a mixture consisting of $N_{\rm I}-1$ noninteracting ions and
$Z_{\rm I}(N_{\rm I}-1)+Z_{\rm A}$ noninteracting electrons: here, each ion
is assumed to have $Z_{\rm B}$ bound-electrons with a charge distribution
$\rho_{\rm b}(r)$ around it and an ionic charge $Z_{\rm I}\equiv
Z_{\rm A}-Z_{\rm B}$. With use of this reference system, the DF theory
can provide an effective external potential $v_{\rm eN}^{\rm eff}(r)$
for electrons around the fixed nucleus. Then, the electron-density
distribution $n_{\rm e}(r|{\rm N})$ around the fixed nucleus
is obtained by solving
the wave equation for $v_{\rm eN}^{\rm eff}(r)$ in the sum of
the bound- and free-electron parts:
\begin{equation}
n_{\rm e}(r|{\rm N})=n_{\rm e}^{0\rm b}(r|v_{\rm eN}^{\rm eff})
=n_{\rm e}^{\rm b}(r|{\rm N})+n_{\rm e}^{\rm f}(r|{\rm N})\;. \label{eq:neN}
\end{equation}
Hence, the bound-electron distribution $n_{\rm e}^{\rm b}(r|{\rm N})$
thus determined constitutes the definition of the ``ion" in
the electron-ion model.
Furthermore, this bound-electron distribution $n_{\rm e}^{\rm b}(r|{\rm N})$
should be taken identical with
the electron distribution $\rho_{\rm b}(r)$ of an ion in the
reference system, since the ion formed around the central nucleus is
necessary to be the same structure as any ion in the system.
Thus, we obtain a self-consistent condition to determine
the distribution $\rho_{\rm b}(r)$ in the premise:
\begin{equation}
\rho_{\rm b}(r)=n_{\rm e}^{\rm b}(r|{\rm N})\equiv n_{\rm e}^{\rm b}(r)
\end{equation}
with the bound-electron number $Z_{\rm B}=\int\rho_{\rm b}(r)d{\bf r}$.
On the other hand, the free-electron part $n_{\rm e}^{\rm f}(r|{\rm N})$
in Eq.~(\ref{eq:neN}) is taken as the electron-ion RDF
$n_0^{\rm e}g_{\scriptscriptstyle\rm eI}(r)$ of the electron-ion mixture
with the free-electron density
$n_0^{\rm e}=Z_{\rm I}n_0^{\rm I}$, and the nucleus-nucleus RDF becomes
the ion-ion RDF $g_{\scriptscriptstyle\rm II}(r)$.
With use of this reference system, we can obtain a tractable expression of
$v_{\rm eN}^{\rm eff}(r)$ for the wave equation to determine
$n_e(r|N)$ by introducing some approximations
to the exchange-correlation term involved in it\cite{NEmodel}:
\begin{equation}\label{eq:vefeN}
v_{\rm eN}^{\rm eff}(r)={\tilde v}_{\rm eI}(r)
-{1\over\beta}\sum_l
\int C_{{\rm e}l}(|{\bf r}-{\bf r}'|)n_0^l
[g_{{\scriptscriptstyle l}{\rm I}}(r')-1]d{\bf r}'\;,
\end{equation}
where $\mu_{\rm XC}(n)$ is the exchange-correlation
potential in the local-density approximation (LDA).
Note that this expression is equal to Eq.~(\ref{eq:Ueff})
without the electron-ion bridge functions $B_{\rm eI}(r)$
except that the bare electron-ion interaction
is explicitly given by
\begin{equation}\label{eq:barveI}
{\tilde v}_{\rm eI}(r)\equiv -{Z_{\rm A}e^2\over r}
+\int v_{\rm ee}(|{\bf r}-{\bf r}'|)n_{\rm e}^{\rm b}(r')d{\bf r}'
+\mu_{\rm XC}(n_{\rm e}^{\rm b}(r)+n_0^{\rm e})
-\mu_{\rm XC}(n_0^{\rm e})\;.
\end{equation}
In this way, the treatment of a liquid metal as a nucleus-ion mixture
is shown to provide the ion-electron model, where the bare electron-ion
interaction ${\tilde v}_{\rm eI}(r)$ and the ionic structure
$\rho_{\rm b}(r)$ can be determined in a self-consistent manner.
With the help of the result from the nucleus-electron model, we can
derive a closed set of integral equations for the ion-electron mixture,
if we introduce the following approximations:
(A) the electron-ion
bridge function in Eq.~(\ref{eq:QHNCei}) is neglected:
$B_{\rm eI}(r)\simeq 0$, (B) the electron-electron DCF $C_{\rm ee}(r)$
is approximated \cite{QHNC} as
\begin{equation}\label{eq:Cee}
C_{\rm ee}(Q)=-\beta v_{\rm ee}(Q)\left[1-G^{\rm jell}(Q)\right]\;.
\end{equation}
using the local-field correction (LFC) $G^{\rm jell}(Q)$ of the jellium model
for an electron gas,
(C) the bare ion-ion potential $v_{\rm II}(r)$ is taken as pure
Coulombic, i.e., $v_{\rm II}(r)=(Z_{\rm I}e)^2/r$,
and (D) the bare electron-ion potential is given by
$v_{\rm eI}(r)={\tilde v}_{\rm eI}(r)$ of Eq.~(\ref{eq:barveI}).
We have called this set of equations the quantal hyper-netted chain
equations because of the approximation $B_{\rm eI}(r)\simeq 0$
in Eq.~(\ref{eq:Ueff}).
\section{MOLECULAR DYNAMICS SIMULATION
BASED ON QUANTAL HYPER-NETTED CHAIN THEORY}
\label{sec:md}
It is important to realize that the electron-ion model leads to
the neutral-fluid model, where the ionic behavior of a liquid metal is
taken to be the same as a neutral
one-component fluid interacting via a binary effective
interaction in treating the ion-ion RDF.
This neutral-fluid model is derived from the electron-ion model, when
an effective ion-ion potential is defined in such a
way that the RDF of a one-component fluid
should become identical with $g_{\scriptscriptstyle\rm II}(r)$ of
the electron-ion mixture:
\begin{equation}\label{eq:gRone}
g(r)\equiv \exp[-\beta v_{\rm eff}(r)+{\it\Gamma}(r)+B(r)]
= g_{\scriptscriptstyle\rm II}(r)\;,
\end{equation}
with use of the Ornstein-Zernike relation for a neutral one-component fluid
\begin{equation}
g(r)-1=C(r)+{\it\Gamma}(r)\;.
\end{equation}
In the above,
the DCF for the one-component fluid is defined by
$n_0^{\rm I}C(Q)\equiv 1-S_{\rm II}(Q)^{-1}$
and ${\it\Gamma}(r)\equiv
\int C(|{\bf r}-{\bf r}'|)n_0^{\rm I}[g(r')-1]d{\bf r}'$.
Thus, we can write the explicit expression for the effective
ion-ion potential of a liquid metal in the neutral-fluid model as
\begin{equation}\label{eq:veff}
\beta v_{\rm eff}(Q)
\equiv \beta v_{\rm II}(Q)
-{|C_{\rm eI}(Q)|^2n_0^{\rm e}\chi_Q^0\over
1-n_0^{\rm e}C_{\rm ee}(Q)\chi_Q^0}\;,
\end{equation}
by taking the bridge function $B(r)$ to be $B_{\rm II}(r)$ of the
electron-ion mixture. Equation (\ref{eq:veff}) can be
interpreted within the scope of the standard pseudopotential theory
by regarding $C_{\rm eI}(r)$ as the pseudopotential
$w_b(Q)=-C_{\rm eI}(Q)/\beta$. In this way, the ion-electron model
is reduced exactly to the neutral-fluid model with a {\it binary}
effective interaction Eq.~(\ref{eq:veff}), in which
the many-body forces are taken into account in the form of the
linear response expression (\ref{eq:veff}),
since the nonlinear effect in the electron screening is involved
in terms of the electron-ion DCF, which plays the role of a nonlinear
pseudopotential.
By noting the above relations (\ref{eq:gRone})\,--\,(\ref{eq:veff}),
the exact expressions (\ref{eq:gII})\,--\,(\ref{eq:geIOZ}) for the
electron-ion model can be
transformed into a set of integral equations: one is the integral
equation for a one-component fluid with the effective ion-ion
potential $v_{\rm eff}(r)$
\begin{equation}\label{eq:QHNCii}
C(r)=\exp[-\beta v_{\rm eff}(r)
+{\it\Gamma}(r)+B(r)]-1-{\it\Gamma}(r)\;,
\end{equation}
and the other an equation for the effective ion-ion interaction
$v_{\rm eff}(r)$, that is expressed in the form of an integral
equation for the electron-ion DCF $C_{\rm eI}(r)$:
\begin{equation}\label{eq:QHNCei}
\hat B C_{\rm eI}(r)
=n_{\rm e}^0(r|v_{\rm eI}
-{\it\Gamma}_{\rm eI}/\beta-B_{\rm eI}/\beta)/n_0^{\rm e}
-1-\hat B {\it\Gamma}_{\rm eI}(r)\;,
\end{equation}
since the effective interaction $v_{\rm eff}(r)$ is given
in terms of $C_{\rm eI}(r)$ by Eq.~(\ref{eq:veff}).
In contrast with the usual effective
potential in the pseudopotential theory, the effective potential
(\ref{eq:veff}) depends on the ion configuration represented
by the ion-ion RDF $g_{\scriptscriptstyle\rm II}(r)$ through the term:
${\it\Gamma}_{\rm eI}(r) \equiv \sum_l
\int C_{el}(|{\bf r}-{\bf r}'|)n_0^l
[g_{{\scriptscriptstyle l}{\scriptscriptstyle\rm I}}(r)-1]d{\bf r}'$
in Eq.~(\ref{eq:QHNCei}).
Under the assumptions (A)\,--\,(D) mentioned before, the QHNC
equations (\ref{eq:QHNCii}) and (\ref{eq:QHNCei}) enables us to
perform an {\it ab initio} molecular-dynamics (MD) simulation
which requires only the atomic number $Z_{\rm A}$ and thermodynamic
states as input parameter, in principle.
The first estimation for $v_{\rm eff}(r)$ can be
obtained with use of $C_{\rm eI}(r)$ evaluated by Eq.~(\ref{eq:QHNCei})
with an initial guess for $g_{\scriptscriptstyle\rm II}(r)$.
Next, an integral equation (\ref{eq:QHNCii}) for a one-component fluid
can be solved by performing the classical MD simulation for this
$v_{\rm eff}(r)$ to produce new ion-ion RDF $g_{\scriptscriptstyle\rm II}(r)$; this
is used again in Eq.~(\ref{eq:QHNCei}) to determine a
new estimation for $v_{\rm eff}(r)$.
This process will be continued until convergence of the effective
ion-ion potential is achieved (we refer to this procedure
as the QHNC-MD method). However, such a straight-forward
repetition of the MD simulation to solve the QHNC equations is not
practical in the viewpoint of the computational cost. Since the
dependence of the effective ion-ion potential $v_{\rm eff}(r)$
on the ionic configuration is rather weak in a simple metal
as we have shown in \cite{QHAl}, we can adopt an approximate theory for
$B(r)$ in Eq.~(\ref{eq:QHNCii}) to get an initial $v_{\rm eff}(r)$
for the QHNC-MD method. For this
purpose, we take the variational modified HNC (VMHNC) equation
proposed by Rosenfeld
\cite{VMHNC}, in which the bridge function is approximated by
$B_{\rm PY}(r;\eta)$ of the Percus-Yevick equation for hard-spheres
of diameter $\sigma$ with the packing fraction $\eta=\pi n_0^{\rm
I}\sigma^3/6$. In the VMHNC equation, the adjustable parameter
$\eta$ is determined by the following condition:
\begin{equation}
{1\over 2}n_0^{\rm I}\int[g(r)-g_{\rm PY}(r;\eta)]
{\partial B_{\rm PY}(r;\eta)\over\partial\eta}d{\bf r}
+{2\eta^2\over(1-\eta)^3}=0\;,
\end{equation}
where $g_{\rm PY}(r;\eta)$ is the RDF for the hard-sphere fluid
with the Percus-Yevick equation. Thus, in a similar way to the QHNC-MD
method, the integral equation (\ref{eq:QHNCii}) in the VMHNC
approximation is solved in a coupled manner with Eq.~(\ref{eq:QHNCei})
producing new effective ion-ion interaction [referred to as the
QHNC-VM method].
Furthermore, an initial potential $v_{\rm eff}(r)$
to this QHNC-VM method can be obtained by
approximating $g_{\scriptscriptstyle\rm II}(r)$ in Eq.~(\ref{eq:QHNCei})
by the step function $\theta(r-a)$ with the ion-sphere radius $a=(4\pi
n_0^{\rm I}/3)^{-1/3}$.
When this final $v_{\rm eff}(r)$ from the QHNC-VM method is used as
an input to the QHNC-MD method, the convergent result can be obtained
by a few repetition of MD simulation.
Finally our procedure to solve
the QHNC equation with the MD simulation (the QHNC-MD method)
is summarized as the flow chart shown in Figure \ref{fig:flow}.
For an initial potential $v_{\rm eff}(r)$ given by
approximation $g_{\scriptscriptstyle\rm II}(r)=\theta(r-a)$,
the QHNC-VM method
in the {\em preparation phase} yields a good initial
guess for the QHNC-MD method.
Then the MD simulation is repeatedly performed to
achieve convergence of $v_{\rm eff}(r)$ in the {\em refinement
phase}.
There are two important points to be noticed regarding the MD simulation
when applied to the QHNC-MD method. One is that the computer simulation
provides the RDF $g(r)$ only within the half of the side length $L$ of
the simulation cell. This causes an unavoidable truncation error
in calculation of the Fourier transform ${\cal F}_Q[g(r)-1]$ to be used
in the evaluation of Eqs. (\ref{eq:veff}) and (\ref{eq:QHNCei}).
The second point is that the MD simulation is performed inevitably on
a truncated potential for a liquid metal whose effective
ion-ion potential is accompanied by a long-ranged oscillatory tail:
the computer simulation may yield different RDF's depending on the
cutoff radius $R_{\rm c}$ of the potential.
Recently we have proposed a precise procedure\cite{BR} to improve
these two defects at the same time and to get the RDF in the whole range
of distance for the full potential $v_{\rm eff}(r)$. This method can
be applied even to the {\it small-size} simulation result for the
truncated potential $u_{\rm c}(r)$:
\begin{equation}
u_{\rm c}(r)\equiv \left\{
\begin{array}{ll}
v_{\rm eff}(r)-v_{\rm eff}(R_{\rm c}) & \mbox{for}\ r<R_{\rm c} \\
0 & \mbox{for}\ r\ge R_{\rm c}
\end{array}
\right.\;.
\end{equation}
As the first step of this procedure, we extract the bridge function
from the raw MD RDF data. For this purpose, we extend the the raw RDF
data of the MD simulation $g_{\rm MD}(r)$, by solving an integral
equation
\begin{equation}\label{eq:grex}
g(r)\equiv
\left\{
\begin{array}{ll}
g_{\rm MD}(r) & \mbox{for}\ r<R \\
\exp[-\beta u_{\rm c}(r)+{\it\Gamma}(r)] & \mbox{for}\ r\ge R
\end{array}
\right.\;,
\end{equation}
coupled with the Ornstein-Zernike relation, where $R$ is the
extrapolating distance ($R<L/2$). At this stage, in order to
obtain a reliable bridge function, it is essential to take $R$ as
short as about 3 to 4 interatomic spacings or simply as $R=R_{\rm
c}$ \cite{BR} and to discard the RDF data outside the distance $R$ so as
to reduce the statistical noise contained in the raw RDF data.
Using the extended $g(r)$, the bridge function $B_{\rm MD}(r)$ can
be extracted for distances where $g_{\rm MD}(r)\ne 0$ by
\begin{equation}\label{eq:MDBr}
B_{\rm MD}(r)\equiv
\left\{
\begin{array}{ll}
\beta u_{\rm c}(r)+\ln[g_{\rm MD}(r)]-{\it\Gamma}(r)
& \mbox{for}\ r<R \\
0 & \mbox{for}\ r\ge R
\end{array}
\right.\;.
\end{equation}
At the second step to get the RDF for the full potential
$v_{\rm eff}(r)$, we solve the integral equation (\ref{eq:QHNCii})
for the full potential with use of this $B_{\rm MD}(r)$ as
an approximation to that of the full potential: this can be
justified by the fact that the bridge
function is not sensitive to the long-range part of the potential
and becomes very weak for long-range distance \cite{BR}.
\section{APPLICATION TO LIQUID ALKALI METALS}
\label{sec:appl}
\subsection{Numerical procedure of QHNC-MD method}
Liquid alkali metals constitute ``simple" metals in the sense that
the bound electrons forming an ion are clearly distinguished from
the conduction electrons and the overlap of the core electrons are
negligible; the approximations (A)--(D) used in the QHNC-MD method
become quit good ones for these metals. Therefore,
we have applied the QHNC-MD method for five liquid alkali metals
(Li, Na, K, Rb, and Cs) near the melting point using the
parameters specified in Table
\ref{tab:param}; the temperature and density have been chosen
to be compared with experimental data
in \cite{ExpLi,ExpNaKCs,ExpNaK,ExpRbCH,ExpRb,ExpRbW}.
Here, the temperature and density of
alkali liquids are specified by two dimensionless parameters:
the plasma parameter $\Gamma=\beta e^2/a$ and
$r_{\rm s}=a/a_{\scriptscriptstyle\rm B}$
in units of the Bohr radius $a_{\scriptscriptstyle\rm B}$,
with the average ion-sphere radius $a$.
In our application of the QHNC-MD simulation to the alkali liquids,
the local-field correction of the jellium model in Eq.~(\ref{eq:Cee})
is chosen to be that proposed by Geldart and Vosko \cite{GV}, since it has
a simple structure and gives a good approximation.
In Eq.~(\ref{eq:barveI}), the expression given by Gunnarsson
and Lundqvist \cite{GL} is adopted as the LDA for the
exchange-correlation potential $\mu_{\rm XC}(n)$.
After the preparation of initial effective potential by
the QHNC-VM method, two iterations
in the refinement phase of the
QHNC-MD method (Figure \ref{fig:flow}) are sufficient to obtain
a convergent solution for alkali liquid metals; 16,000 particles have
been used in the first MD run (Run-1) and 4,000 particles for the
second MD run (Run-2). The MD simulation has been performed over
50,000 time steps for Run-1 and 100,000 time steps for Run-2
with the cubic periodic boundary conditions;
the temperature of the system is kept constant by the isokinetic
constraint \cite{Hoover}. The
equations of motion are integrated by a fifth order differential
algorithm \cite{Algorithm} with the time increment
${\it\Delta}t=0.0025n_0^{{\rm I}\;-1/3}(m_{\scriptscriptstyle\rm I}\beta)$
with the mass of an ion $m_{\scriptscriptstyle\rm I}$:
the corresponding real time is shown in Table \ref{tab:param}.
In each MD simulation of Run-1 and Run-2 for
iterations, the effective potential is
cut at the radius $R_{\rm c}$ located at the node of
the Friedel oscillation of $v_{\rm eff}(r)$ as shown in
Table \ref{tab:param}. All MD simulations have been
carried out on a vector-parallel processor Monte-4 \cite{Monte4} at
Japan Atomic Energy
Research Institute. The computational time required for 10,000
steps is about 30 to 50 hours for 16,000 particles including the
sampling of the RDF.
The integral equation (\ref{eq:grex}) has been solved by an
iterative procedure introduced by Ng \cite{Ng} to extend the
raw MD data $g_{\scriptscriptstyle\rm MD}(r)$ to the
whole $r$ range: the extending distance $R$ in this
procedure (\ref{eq:grex}) is taken to be
$R_{\rm c}$ for the whole cases. The number of grid points and step size
used in numerical integrations are 1024 points and
${\it\Delta}r=0.025a$, respectively. Using $C(r)$ obtained by the
HNC equation as an initial input function, it takes about 10,000
iterations to achieve convergence.
In order to examine both numerical and computational efficiency of
the QHNC-MD method, we have tested the convergence of the RDF
by evaluating following consistency
measure for $g(r)$:
\begin{eqnarray}
{\it\Delta}g_i(r) & \equiv & g_i(r)-g_{i-1}(r)\;,\label{eq:diff} \\
|{\it\Delta}g_i| & \equiv &
\left({4\pi n_0^{\rm I}\over
3}\int_0^\infty|{\it\Delta}g_i(r)|^2r^2dr\right)^{1/2}\;.
\label{eq:measure}
\end{eqnarray}
Here $g_i(r)$ is $g_{\scriptscriptstyle\rm II}(r)$ obtained by the
$i$-th MD simulation and $g_0(r)$ is that obtained by the final
step of the preparation phase.
Figure \ref{fig:gconv} shows the consistency
measure (\ref{eq:diff}) of the present QHNC-MD simulation for
the case of liquid Li, yielding the values $|{\it\Delta}g_1|
=1.53\times 10^{-1}$ and $|{\it\Delta}g_2|=5.92\times 10^{-3}$.
It is easily seen that the
convergence of $g_{\scriptscriptstyle\rm II}(r)$ is very fast;
the difference of $g_{\scriptscriptstyle\rm II}(r)$
between Run-1 and Run-2 is
situated almost within the statistical error of the sampling of the
RDF in the MD simulation: this means accuracy of about 3 to 4
digits is already achieved in the QHNC-MD calculation of Run-1. In
a consistent way to the convergence of the RDF, a good convergence
of the effective ion-ion potential $v_{\rm eff}(r)$ is achieved
in Run-1.
Similar to the case of liquid Li, a good
convergence of $g_{\scriptscriptstyle\rm II}(r)$ and $v_{\rm eff}(r)$
is also achieved for other liquid metals.
It should be noted that the preparation phase of the QHNC-MD
method largely enhances the convergence of $v_{\rm eff}(r)$.
Concerning the treatment of the raw MD data for the ion-ion RDF, it
should be emphasized that the extension procedure\cite{BR} to obtain the RDF
for the full potential applied to the raw MD data is indispensable
for the present calculation. This situation is illustrated in
Fig.~\ref{fig:Ligrdiff}, where the truncation error in
$g_{\scriptscriptstyle\rm II}(r)$ due to
the use of the cutoff potential in the MD simulation is so large
that the convergence of the QHNC-MD method will not be attained
with the raw RDF data. In addition, the extended RDF for the full
potential is almost identical with the result of VMHNC equation for
the same potential (Figure \ref{fig:Ligrdiff}) for $r\agt 5a$,
i.e., after the third peak of RDF. This suggests that the long
ranged Friedel oscillation of $v_{\rm eff}(r)$ typically seen for
liquid metals is essential for the detailed structure of the RDF at
long distances, and it is necessary to include the information of
the long-range part of $v_{\rm eff}(r)$ into
$g_{\scriptscriptstyle\rm II}(r)$ in
order to obtain a self-consistent solution of the effective
ion-ion potential by the QHNC-MD calculation.
It is concluded that the convergence of the present QHNC-MD
simulation is well attained from both numerical and computational
points of view, helped by a good initial
estimation from the VMHNC equation in the
preparation phase; the extension procedure to obtain the full-range RDF
for the uncut effective potential is also necessary to get the convergence.
\subsection{Ion-ion and electron-ion radial distribution functions}
Following the procedure mentioned above, we obtain the effective interatomic
interaction, the ion-ion and electron-ion RDF's, the charge distribution
$\rho(r)$ of neutral pseudoatom,
the bound electron distribution $n_{\rm b}(r)$ forming an ion and the
bridge functions for liquid alkali metals in a self-consistent way from the
atomic number as the only input. In the first place, we show
structure factors, the Fourier transform of the ion-ion RDF's. It is importa
nt
for a detail comparison with experiment to use the MD RDF corrected for a
full potential and extrapolated to whole range of distance
in its Fourier transform.
Figure \ref{fig:LiSQ} exhibits the structure
factors for liquid Li calculated by the QHNC-MD simulation
in comparison with the
experimental result\cite{ExpLi}.
It is clearly seen that the present
result is in excellent agreement with the experiment,
improving the detailed structure of $S_{\rm II}(Q)$ near its second
peak compared with the result of the VMHNC equation (the final
result of the preparation phase). This improvement on $S_{\rm
II}(Q)$ by the refinement phase essentially relies on the detailed
oscillatory behavior of the bridge function extracted from the raw
RDF of the MD simulation. The discrepancy in the bridge function
between the MD simulation and VMHNC equation gives no serious effects
on the first peak of the RDF. But details of the RDF for $2a\alt r\alt 5a$
are rather sensitive to the oscillation of $B(r)$ in a similar way as
discussed in \cite{BR}. Therefore the use of the extracted bridge
function $B_{\scriptscriptstyle\rm MD}(r)$ to determine the corrected RDF is
important in order to guarantee the correct behavior of the
structure factor by the Fourier transform.
The structure factors calculated for Na and K are shown in Fig.~5
in comparison with the neutron and X-ray experiments\cite{ExpNaKCs,ExpNaK}.
The QHNC-MD structure factor of Na is in excellent agreement with the X-ray
result (full circles) showing a small deviation
from the neutron data (circles) at the first peak,
while the curves of the QHNC-MD structure factor for K lies
between the neutron (circles) and X-ray (full circles) experimental results.
Also, the structure factors from the QHNC-MD method for Rb and Cs are
compared with the neutron and X-ray
experiments\cite{ExpRbCH,ExpRb,ExpRbW} in Fig.~6.
The first peak of Rb structure factor observed by neutron
experiment (full circles)\cite{ExpRbCH} is shifted a little to large $Q$
side compared with that observed the X-ray experiment
(circles)\cite{ExpRbW}; the QHNC-MD result
has the same first-peak position to the X-ray data with a little different
height and shows overall agreement with the neutron observation.
On the other hand, the QHNC-MD structure factor of Cs becomes
higher in the first-peak
than the experiment\cite{ExpNaKCs} and shows a small
deviation in the phase of oscillation
of structure factor for the large $Q$ region compared with experiment.
In our treatment, we solve the Schr\"{o}dinger equation to
obtain the bound-electron distribution and electron-ion RDF;
the relativistic effect is not taken
into account. In the case of Cs, the relativistic effect may
contribute to the
calculation of the structure factor, since its atomic number
$Z_{\rm A}=55$ is rather large; there is a possibility
that this effect may be ascribed to
this small discrepancy between the calculated and experimental results in
Cs structure factor. In the same way as shown in the Li structure factor,
the QHNC-VM structure factors deviate from the QHNC-MD results near the their
second peak for Na, K, Rb and Cs as seen Figs. 5 and 6.
This second-peak difference between the QHNC-MD and QHNC-VM structure factors
brings about a distinct difference in the RDF's obtained from their Fourier
transform, as can be seen from an example of Na shown in Fig.~7.
Thus, the RDF from the VMHNC equation is found not so exact as to compare
with details of experimental results for alkali liquids; the QHNC-MD method
is shown to produce reliable results for all alkali liquids.
Next, we proceed to discuss the electronic structure around ion.
The electron-ion RDF's from the QHNC-MD method are shown in
Fig.~8 for the case of Rb at temperature 313 K together with the ion-ion RDF.
The electron-ion structure factor $S_{\rm eI}(Q)$ is represented in the
form\cite{hyd}:
\begin{equation}\label{eq:sei2}
S_{\rm eI}(Q)= {\rho(Q) \over \sqrt{Z_{\rm I}} }S_{\rm II}(Q)\,,
\end{equation}
in terms of the ion-ion structure factor and the charge distribution of the
pseudoatom:
\begin{equation}\label{eq:rhoQ}
\rho(Q)\equiv {n_0^e C_{\rm eI}(Q)\chi_Q^0 \over 1-n_0^e
C_{\rm ee}(Q)\chi_Q^0 }\,\,. \\
\end{equation}
With the help of the above equations, the pseudopotential method using
the Ashcroft model potential can evaluate the electron-ion RDF by
inserting it in (\ref{eq:rhoQ}) instead of $-C_{\rm eI}(r)/\beta$;
its result using the empty core radius 1.27${\rm \AA}$ has no inner-core
structure near the origin as is expected.
In the QHNC-MD method, we need not introduce any pseudization
in treating the core region. Therefore, the electron-ion
RDF $g_{\scriptscriptstyle\rm eI}(r)$ exhibits the inner-core
structure similar to a free-atom as is shown in
Fig.~8 in contrast with result of the pseudopotential method (full circles).
It is interesting to note that the electron-ion RDF has an oscillation
with an inverse phase to that of the ion-ion RDF,
since the electrons are pushed
away by an ion as a whole in the core region. The charge distribution
$\rho(r)$ of the pseudoatom in a liquid Rb has a similar structure
to the distribution $\rho_{\rm 5s}(r)$ of the $5s$-electrons in
the free-atom; therefore, the total electron-density distribution
$n_{\rm b}(r)+\rho(r)$ is almost same to that of a free atom as was
indicated by Ziman.
Also, note that the positions of the dips in the electron-ion RDF are
coincident with those appeared in the $5s-$electron charge
distribution $\rho_{\rm 5s}(r)$ in
a free-atom; these dips in the electron-ion RDF reflect
the inner structure of an ion in a metal.
The electron-ion RDF's of liquid alkali metals (Li, Na, K, Rb and Cs)
are shown in Fig.~10, where the inner-core structure is omitted near
the origin in each curve.
The dip in the electron-ion RDF becomes shallower and is shifted to the
right side indicating the growth of the core size, as the atomic number
increases from Na to Cs: the curve of Li shows an exception to this tendency.
When the electron-ion RDF is determined by solving the wave equation for the
self-consistent potential $v_{\rm eI}^{\rm eff}(r)$, we obtain from
Eq.~(\ref{eq:QHNCei}) the electron-ion DCF $C_{\rm eI}(r)$, which yields
an effective ion-ion potential $v_{\rm eff}(r)$ of
Eq.~(\ref{eq:veff}). Corresponding to each
curves of $g_{\rm eI}(r)$ in Fig.~10, the effective ion-ion
potential $v_{\rm eff}(r)$ is determined for each element
as shown in Fig.~11. It should be
emphasized that there is no units for scaling lengths and energies to
effective ion-ion potentials determined by the QHNC-MD method in contrast
with those obtained by using the Ashcroft model potential\cite{Ba92}.
Similarly, no scaling features are found in the effective ion-ion potential
obtained by the Dagens-Rasolt-Taylor method\cite{DRT},
which can be thought of as an approximation to the QHNC
formulation in the sense that the ion-ion RDF is there
replaced by the spherical vacancy in Eq.~(\ref{eq:vefeN})\cite{DRTcal}.
Nevertheless, it is shown that for liquid alkali metals near the melting
point the ion-ion RDF's from the QHNC-MD method can be
scaled almost into one curve by taking the average ion
radius $a$ in the units of length as indicated in Fig.~12 except Cs.
As the summarization of this section, we conclude that the present
application of the QHNC-MD method to liquid alkali metals provides
the ionic structures in excellent agreement
with the experiments within the computational capacity
available at present, enabling to handle a relatively large system
size in the MD simulation in contrast with the usual {\it ab
initio} simulation for the electron-ion mixture. So as to
be consistent with the ionic structure specified
by the ion-ion RDF, the QHNC-MD method is shown to give the electron-ion
RDF $g_{\scriptscriptstyle\rm eI}(r)$,
the charge distribution $\rho(r)$ of a pseudoatom and the density
distribution $n_{\rm b}(r)$ of the bound electrons forming an ion in
a liquid metal at the same time; alternatively
this electronic structure produces the
ion-ion potential $v_{\rm eff}(r)$ used for the MD simulation.
\section{DISCUSSION}
\label{sec:disc}
We have shown that the QHNC-MD formulation provides a very precise
description of simple liquid metals at any state
from the atomic number $Z_{\rm A}$ as the only input: this method is
proven to yield a first-principles calculation. Our previous
calculations with the VMHNC equation, which has been used for input
to the QHNC-MD method, generated the structure factors with
a small but systematic deviation from experiments near the second peak.
Now our QHNC-MD method is shown to correct this deviation and
to yield results in excellent agreement with
experiments in the whole range of wavenumber $Q$.
Even in alkali metals, there is no scaling unit of lengths and energies
for the effective ion-ion potentials $v_{\rm eff}(r)$ determined from the
QHNC-MD method; each potential is different from other and reflects a
difference in the bound-electron structure of an ion in each metal.
The situation is
different in the case of the effective potential calculated by use of
the Ashcroft pseudopotential, which leads to give almost single potential
curve for all alkali liquids if scaled
with the proper unit\cite{Ba92}. However, it is interesting to note
that the ion-ion RDF's can be scaled by the unit of $a$, the average
ion-sphere radius, which enables plot the results in almost single
curve for alkali liquids (Li, Na, K, Rb) near the melting point
except the case
of Cs. This fact was already found in the experimental results for
the structure factors\cite{ExpNaKCs}.
In our QHNC-MD formulation, the exchange-correlation effect expressed by
the LFC $G(Q)$ and the LDA $\mu_{\rm XC}(n)$ are introduced from outside the
framework of our formulation: those of the jellium model, where the ion
distribution is replaced by the positive uniform background. The
jellium model gives a good description for the electrons in alkali
metals; the present success of the QHNC-MD method depends on this fact. In
addition, the structure factors of alkali metals calculated by this
method are almost independent of what kind of LFC to choose.
However, it should be noted that the LFC in the QHNC formulation
should depend on the ion configuration precisely, since it is
defined for the electron-electron DCF in the electron-ion mixture.
To compare the MD structure factors with experiments in detail, it
is important to extrapolate the MD RDF to large distances and to correct
errors caused by the cut effective ion-ion potential used in the
MD simulation. Our extrapolation
method\cite{BR} is shown very efficient in dealing with the raw
MD data for a liquid
metal with ion-ion potential accompanied by a long-range
Friedel oscillation; this extrapolation method is indispensable
to obtain a convergent solution in the QHNC-MD method.
In order to obtain so reliable bridge function
for the extrapolation of the MD RDF, it is necessary to
get a reasonable statistical accuracy for evaluation of the RDF;
for this purpose, the MD simulation must be performed for at
least several thousand particles taking
about $10^{10}$ to $10^{11}$ samples
\cite{BR}.
The Car-Parrinello MD (CP-MD) method is based on the same ground to
the QHNC-MD method: the electron-ion mixture model for liquid metals and the
jellium model for electrons, which are treated by the
DF theory. Also, the bare ion-ion interaction is taken
as a pure Coulombic in both treatment. However, in the CP-MD method,
the bare electron-ion interaction is approximated by a
pseudopotential, which is introduced from outside of the
CP-MD formulation; this fact makes a contrast with our QHNC-MD
method where it is obtained self-consistently within the framework
of the QHNC formulation and no pseudization is necessary
in treating the core region. As a consequence, the electron-ion RDF
extracted from the CP-MD result does not have an inner core structure.
It should be emphasized that
in the QHNC formulation the inner electronic structure around
a fixed nucleus, such as
the electron-ion RDF in the core region and the bound-electron
distribution $n_{\rm b}(r)$ of an ion in a liquid metal, is
determined to be
consistent with the outer structure, that is, with the surrounding
ion and electron configurations.
Therefore, the QHNC-MD method can be applied to a high density plasma,
where a usual pseudopotential can not be constructed due to the fact
the atomic structure in a highly compressed plasma state is quite
different from that in the vacuum.
While in the CP-MD treatment
the electron distribution is determined for the multi-ion
configuration and the ions are considered to be interacting via
many-body forces, our approach deals with the one-center problem
to determine the electron and ion distributions around a fixed ion
in a liquid metal.
As this connection, it should be kept in mind that the
electron-ion mixture can be exactly taken as a quasi one-component
system interacting only via a {\it pairwise} interaction in
the evaluation of the ion-ion RDF for simple liquid metals.
The advantage of the present method against the CP-MD method based on the
usual pseudopotential theory can be summarized as follows: (1) The
present procedure is capable to handle a large system size ($\sim$
$10^3$ to $10^4$ particles) in the MD simulation within the
computational resources available at present, helped by the good
initial guess with the VMHNC approximation for solving the QHNC
equation. (2) In the QHNC-MD method, the many-body forces and
nonlinear effect in the
electron screening are taken into account automatically in the form
of a pairwise interaction in such a way that the nonlinear
pseudopotential is constructed in terms of $C_{\rm eI}(r)$. (3) By
setting up an additional integral equation for $C_{\rm ee}(r)$, our
method can treat the case where the jellium model for the electrons
in a metal breaks down, that is, where the exchange-correlation effect
begins to depend on the the ion configuration\cite{hyd}.
Therefore, (4) our method is applicable to high density plasmas
where the ionic structure becomes significantly so different from a
free atom due to high compression that the usual pseudopotential
theory cannot be applied.
The QHNC-MD method is developed on the base of the ion-electron model,
where the bound electrons are assumed to be clearly distinguished
from the conduction electrons and
the ions are so rigid and so small that the ion-ion bare
interaction is taken as a pure Coulombic
$v_{\scriptscriptstyle\rm II}(r)=(Z_{\rm I}e)^2/r$.
Therefore, our method is applicable only to a simple metallic system.
In a transition metal, for example, an ``ion" cannot be clearly
defined since the bound electrons are not distinct from
the conduction electron, and the overlap of ``ions" is significant:
many-body interactions becomes important. Our method cannot be
applied to such a case. While the CP-MD method treat electrons in the
multi-center problem, our method treat them as the single-center
problem to determine the density distribution around an fixed ion
in a liquid metal. Thus, our method cannot describe states of the
electrons in a multi-center configuration, such as the density
of states. Moreover, it should be noticed that an orbital of an
ion determined by the QHNC-MD simulation may be taken as an average
orbital in some sense compared with that determined the CP-MD
simulation: there, many-body forces are changing at every
time-step, while the binary ion-ion interaction remains constant
at every time-step in the QHNC-MD simulation.
Already, we have proposed a method to treat non-simple
metals\cite{QHNC}; the RDF's, the ionic charge and
the muffin-tin potential are determined by solving a single-center
problem with a bare ion-ion interaction as an input,
while the density of states, the bare ion-ion interaction and
the thermodynamic properties are to be obtained as results of
the multi-center problem with use of output from the single-center problem.
The CP-MD method can produce the electron-ion RDF and
the electron-ion structure $S_{\rm eI}(Q)$ is obtained from
its Fourier transform. Thus, the charge distribution $\rho(Q)$ of
a neutral pseudoatom can be calculated from Eq.~(\ref{eq:rhoQ})
even in the CP-MD
simulation, since the relation (\ref{eq:sei2}) between $\rho(Q)$
and $S_{\rm eI}(Q)$ is exact if a liquid metal can be treated as a
ion-electron mixture. Also, the electron-ion DCF $C_{\rm eI}(Q)$ is
determined from Eq.~(\ref{eq:rhoQ}) to give an effective ion-ion
interaction $v_{\rm eff}(r)$ from Eq.~(\ref{eq:veff}); these
quantities must be consistent
to the ion-ion structure factor $S_{\rm II}(Q)$, if the LFC $G(Q)$
in the electron-electron DCF $C_{\rm ee}(r)$ is chosen exact one in
the electron-ion mixture. In this connection, there is a point to
notice regarding the CP-MD method that the exchange-correlation
effect of the valence electrons is treated by the LDA approximation:
the LFC $G(Q)$ does not appear in the CP-MD method. This consistency
test in the CP-MD method can be exemplified by applying it to
liquid alkali metals, which are typical examples of the
electron-ion model. At the same time, these quantities in the CP-MD
method can be compared with the results of the QHNC-MD method, both methods
providing first principles calculations.
\section*{ACKNOWLEDGMENTS}
We would like to thank Center of promotion of computational
science and engineering, Japan Atomic Energy Research Institute
for permitting us to use a
plenty amount of computational resources on the dedicated
vector-parallel processor Monte-4.\\
\vspace{.2cm}
It is a sad duty to inform the scientific community that Dr. Shaw Kambayashi
died on April 3, 1995 following a car accident
just after his thirtieth birthday.
|
\section{Introduction}
$\qq$ In the recent analysis of two dimensional(2d) quantum gravity(QG), the
conformal approach or the matrix model approach have been intensively
done. Those approaches are nonperturbative ones and
it is expected that some non-perturbative features are
important to understand the theory. At the same time, however,
it is known that an orthodox perturbative approach, the semiclassical
approach, is also useful in 2d QG\cite{DJ,SEI}.
We present a close examination of the latter approach.
A key point in this treatment is how to treat the area constraint
and the topology constraint. The regularization of infrared divergence
(and ultraviolet divergence in the quantum evaluation) is also important.
The semiclassical treatments of 2d QG so far are insufficient in these
points. We present a new formalism.
Despite of the long period of research, the physical picture of the
Liouville theory, in relation to 2d QG, seems obscure.
It shows the delicacy or the subtlety
of the theory and requires some other proper
formalism and regularization.
As far as the popular formalism based on the conformal field theory
is taken,
the barrier $c_m=1$~ does not seem to be overcome.
Whereas the computer simulations seem to no special difficulty for
the prohibitted region of $c_m$\cite{AJT,AT}.
This conflict seems gradually serious.
Although the problem is examined from different approaches,
it is fair to say the true situation is
not known at present. Recently it has been shown that
the semiclassical results nicely explain
the simulation data\cite{ITY1,ITY2}.
It is well known, in the lattice QG,
that a higher-derivative term, $\be R^2$, regularize the theory very well.
It has the effect of smoothing the surface (if we take
a proper coupling sign).
The importance of the term is also stressed in the continuum context\cite{KPZ}.
{}From the simple power-counting we see the ultra-violet
behaviour becomes well regularized.
We can take two standpoints about the $R^2$-term:\
1)\ We are considering the 2d $R^2$-QG as one gravitational model;\
2)\ We regard $R^2$-term as a regularization to define the $\be=0$~theory
and expect its effect
disappears when some limit is taken.
Although 1) is mainly taken in the present paper,
both standpoints are important at this time of development.
In the semiclassical analysis of 2d $R^2$-QG in \cite{ITY1,ITY2}
one of
the classical solutions (positive curvature solution)
is analysed. In the present paper we present
the full structure of the classical vacua.
It is quite interesting that the positive and negative solutions are
symmetric with respect to a reflection in the coupling $\be$-space.
The explanation is self-contained.
It is well known that the global quantities in the gravitational
system, such as entropy, volume and temperature of the total universe, obey
the laws of thermodynamics. It says those quantites can be
regarded as thermodynamic state variables of an equilibrium state. We will find
those properties in the present 2d model of QG. We can characterize all phases
appearing in the theory by the $\be$-dependence of the temperature.
We present a general formalism in Sec.2, where some thermodynamic
functions are introduced. In Sec.3 the classical solutions are obtained.
They are $R^2$-gravity
version of the Liouville solution and describe positive
and negative constant-curvature manifolds. The analytic expressions
of some physical quantites are given and analysed.
We characterize all asymptotic regions of the solutions in Sec.4.
In Sec.5 an integral about
a parameter $\la$, which appears in the general formalism in connection with
the area costraint, is done. The analytic expressions of cross-over
points in the theory are obtained in Sec.6. It shows the essetial
behaviour of the theory is controled by a toatl derivative (global) term.
In Sec.7 we characterize each phase and obtain the equation of state.
The expressions for temperature and entropy are obtained.
We conclude in Sec.8.
\section{Semiclassical Quantization and Thermodynamic Functions}
$\qq$ Before the concrete evaluation, we describe here the present new
formalism.
We take the Euclidean action,
\begin{eqnarray}
& S_{tot}=S_{gra}+S_m\com \q
S_{gra}[g;G,\be,\mu]=\intx\sqg (\frac{1}{G} R-\be R^2-\mu)\com
& \nn\\
& S_m[g,\Phi;c_m]=-\intx\sqg (\half\sum_{i=1}^{c_m}\pl_a\Phi_i\cdot
g^{ab}\cdot \pl_b\Phi_i)\com\q (\ a,b=1,2\ )\com & \label{3.0}
\end{eqnarray}
under the fixed area condition
$ A=\intx \sqg\ $. Here
$G$\ is the gravitaional coupling constant, $\mu$\ is the cosmological
constant ,
$\be$\ is the coupling strength for $R^2$-term and $\Phi$\ is the $c_m$-
components scalar matter fields.
The 2 dim quantum gravity can be treated in the way similar to
the flat theory by
taking the conformal-flat gauge($a,b=1,2$),
\begin{eqnarray}
& g_{ab}=\ e^{\vp}\ \del_{ab}\com
(\del_{ab})=
\left( \begin{array}{cc} 1 & 0 \\
0 & 1
\end{array}\right)
\com& \label{3.1}
\end{eqnarray}
the action~(\ref{3.0})
gives us,after integrating out the matter fields and Faddeev-Popov ghost,
the following partition function\cite{P}.
\begin{eqnarray}
& \int\frac{\Dcal g\Dcal\Phi}{V_{GC}}\{exp\PLinv S_{tot}\}~\del(\intx\sqg-A)
=exp\PLinv (\frac{8\pi(1-h)}{G}-\mu A)\times Z[A]\com & \nn\\
& Z[A]\equiv\int\Dcal\vp~ e^{+\frac{1}{\hbar}
S_0[\vp]}~\del(\intx ~e^\vp - A)\com & \label{3.3}\\
& S_0[\vp]=\intx\ (\frac{1}{2\ga}\vp\pl^2\vp
-\be~e^{-\vp}(\pl^2\vp)^2 +\frac{\xi}{2\ga}\pl_a(\vp\pl_a\vp)\ )\com
\q \frac{1}{\ga}=\frac{1}{48\pi}(26-c_m)\com & \label{3.2}
\end{eqnarray}
where the relations for
Einstein term and the cosmological term:\
$\intx\sqg R =8\pi (1-h),\ h=\mbox{number of handles},
\intx\sqg =A $\ ,are used\cite{N2a}.
$V_{GC}$\ is the gauge volume due to the general coordinate invariance.
$\xi$\ is a free parameter. The total derivative term generally appears when
integrating out the anomaly equation
\ $\del S_{ind}[\vp]/\del\vp=\frac{1}{\ga}\pl^2\vp\ $.
This term turns out to be very important\cite{N2b}.
We consider the manifold of a fixed topology of
the sphere ,$h=0$, and with the finite area $A$.
Furthermore we consider the case $\ga>0\ (c_m<26)$\cite{N2c}.
\ $\hbar$\ is Planck constant\cite{N2d}.
The Laplace transform of (\ref{3.3}) is given by
\begin{eqnarray}
&{\Zhat}[\la]=
\int_0^{\infty} Z[A]e^{-\la A/\hbar}~dA & \nn \\
&\qqq=\int\Dcal\vp~exp[~
+\frac{1}{\hbar}\{ S_0[\vp]-\la\ \intx ~e^\vp \} ]\pr &
\label{3.4a}
\end{eqnarray}
As the arguments of
$Z$,(\ref{3.3}), and $\Zhat$,(\ref{3.4a}),
we do not write explicitly $\be,\ga$-
dependence .
(\ref{3.3}) is the micro canonical distribution for the fixed area $A$\ ,
whereas (\ref{3.4a}) is the grand canonical distribution (variable
area ) with the chemical potential $\la$\ . From (\ref{3.4a}), we obtain
the expectation value for the area as
\begin{eqnarray}
<A_{op}>=\frac{1}{\Zhat}\frac{d}{d(-\la/\hbar )}\Zhat[\la]
\equiv <\intx e^{\vp}>_{\Zhat}\com\q
A_{op}\equiv\intx~e^\vp\pr \label{3.4c}
\end{eqnarray}
By inverting this equation we obtain \ $\la=\labar(<A_{op}>)$\ . Equivalently
we also abtain it in terms of the Legendre-transformed generating
function $\Ga[<A_{op}>]$.
\begin{eqnarray}
\Ga[<A_{op}>]\equiv ln~\Zhat[\labar]+\PLinv\labar\times <A_{op}>\com\ \
\labar(<A_{op}>)=\hbar\frac{d\Ga[<A_{op}>]}{d<A_{op}>}\pr
\label{3.4d}
\end{eqnarray}
$Z[A]$\ can be obtained from $\Zhat[\la]$\ by the inverse Laplace
transformation.
\begin{eqnarray}
& Z[A]=\int d\la~\Zhat[\la]~e^{+\la A/\hbar}
\equiv\int d\la~ Y[A,\la]\com & \nn\\
& Y[A,\la]\equiv
\int\Dcal\vp~exp\ \frac{1}{\hbar}[\ S_0[\vp]
-\la (\intx e^\vp - A)] & \label{3.4f}\\
& \Ga^{eff}[A,\la]\equiv~ln~Y[A,\la]\com\nn
\end{eqnarray}
where $\la$-integral should be carried out along an appropriate contour
parallel to the imaginary axis of the complex $\la$-plane(Fig.1).
{\vs 5}
\begin{center}
Fig.1\q The contour of $\la$-integral in the complex $\la$-plane
\end{center}
$\Zhat[\la]$ is defined
by (\ref{3.4a}). The partition function (\ref{3.4f}) can be evaluated
semiclassically in the following two steps (i) and (ii).
In the evaluation we will
relate above thermodynamic functions.
\flushleft{(i)\ $\vp$-integral}
First we define some quantities.
\begin{eqnarray}
& S_\la[\vp]\equiv S_0[\vp]-\la\intx~e^\vp\ & \nn\\
& =\intx\ (\frac{1}{2\ga}\vp\pl^2\vp -\be~e^{-\vp}(\pl^2\vp)^2\
+\frac{\xi}{2\ga}\pl_a(\vp\pl_a\vp) -\la~e^\vp\ )\com &\nn\\
&\Zhat[\la]=\int\Dcal\vp~exp~\{\PLinv S_\la[\vp]\}
\equiv\ exp~\PLinv\Gahat[\la]&
\label{3.4g}
\end{eqnarray}
$\Gahat(\la)$\ is the effective action corresponding to $S_\la[\vp]$.
$\Gahat(\la)$ can be evaluated loop-wise\cite{N2e}
by the semiclassical expansion.
\begin{eqnarray}
\vp(x)=\vp_c(x;\la)+\sqrt{\hbar}~\psi(x)\com
\label{3.4h}
\end{eqnarray}
where we take the 'mean field' (or 'background field') as the solution
of the classical field equation for $S_\la[\vp]$.
\begin{eqnarray}
\left.\frac{\del}{\del\vp}S_\la[\vp]\right|_{\vp_c}=\ 0\pr
\label{3.4i}
\end{eqnarray}
Then $\Gahat[\la]=\hbar~ln~\Zhat[\la;\vp]$ can be evaluated as
\begin{eqnarray}
&\Zhat[\la]=\int \Dcal\psi~
exp~\PLinv S_\la[\vp_c+\sqrt{\hbar}\psi] &\nn\\
&=~exp~\PLinv S_\la[\vp_c]\times \int \Dcal\psi~
exp\{~\frac{1}{2}\frac{\del^2S_\la}{\del\vp^2}|_{\vp_c}\psi\psi
+O(\sqrt{\hbar}\psi^3) \} &\label{3.4j}\\
&\equiv exp\{~\PLinv\Gahat^0[\la]+\Gahat^1[\la]+
O(\mbox{higher-than 1-loop,\ }\hbar )~\}\com & \nn\\
&\Gahat^0[\la]\equiv S_\la[\vp_c]\com& \nn\\
&\Gahat[\la]=\Gahat^0[\la]+\hbar\Gahat^1[\la]
+O(\mbox{higher-than 1-loop,\ }\hbar^2 )\com&\nn
\end{eqnarray}
where $\Gahat^n[\la],(n\geq 1),$\ is the n-loop quantum effects.
In (\ref{3.4j}),
the effect up to 1-loop order is explicitly written.
\flushleft{(ii)\ $\la$-integral}
Using the result of (i), $Z[A]$ can be written as
\begin{eqnarray}
Z[A]=\int d\la~exp~\PLinv\{~\Gahat[\la]+\la A~\}\pr
\label{3.4k}
\end{eqnarray}
The $\la$-integral of (\ref{3.4k}) can be again evaluated in the
semiclassical way as follows. The dominant value $\la_c$\ is defined by,
\begin{eqnarray}
\frac{d}{d\la}(\Gahat[\la]+\la A)|_{\la_c}
=\frac{d\Gahat[\la_c]}{d\la_c}+A=0\com \nn\\
\Gahat[\la]=\Gahat^0[\la]+\hbar\Gahat^1[\la]+\cdots\com\q
\la_c=\la_c^0+\hbar\la_c^1+\cdots\com\label{3.4l}
\end{eqnarray}
where $\la_c^n$ \ is the n-loop effect of $\psi$-integration and
is recursively obtained.
The $\la$-integral is approximately obtained by
evaluating the fluctuation around $\la_c$\ perturbatively
as follows.
\begin{eqnarray}
& \la=\la_c+~\om\com & \nn\\
& Z[A]
=exp~\PLinv\{\Gahat[\la_c]+\la_c A\}\times\int d\om~
exp~\PLinv\{~\frac{1}{2}
\left.\frac{d^2\Gahat[\la]}{d\la^2}\right|_{\la_c}~\om^2
+O(\om^3) \ \}\pr &\label{3.4ll}
\end{eqnarray}
This integral will be evaluated in Sec.5.
$\qq$ Here we note that the equation (\ref{3.4l}),
by which $\la_c$\ is defined,
is re-written as
\begin{eqnarray}
A=\ -\frac{d\Gahat[\la_c]}{d\la_c}=\left.\frac{1}{\Zhat[\la]}
\frac{d\Zhat[\la]}{d(-\la/\hbar)}\right|_{\la_c}\com
\label{3.4m}
\end{eqnarray}
which is exactly the same as (\ref{3.4c})
by identifying $A$\ above with $<A_{op}>$ \ in (\ref{3.4c}).
Therefore $\la_c(A)=\labar(A)$\ . And
$exp\{\PLinv\Gahat[\la_c]\}=~\Zhat[\la_c]$ ,in
the second equation of(\ref{3.4ll}) ,is exactly
the same quantity as $\Zhat[\labar(A)]$ in (\ref{3.4d}).
Furthermore $\Ga^{eff}[A,\la_c]$\
exactly coincides
with $\Ga[A]$\ of (\ref{3.4d}).
\begin{eqnarray}
ln~Z[A]\approx\Ga^{eff}[A,\la_c]=\PLinv(\Gahat[\la_c]+\la_c A\ )
=\Ga[A]\pr \label{3.4n}
\end{eqnarray}
$\qq$ We have introduced some thermodynamic functions. For their comparison
they are listed in Table 1.
\vspace{0.5cm}
\begin{tabular}{|c|c|c|c|c|}
\hline
Indep. & Integral & Special & Partition Func.
& Effective \\
param. & var. & func. & & action \\
\hline
$A$ & $\vp(x)$ & & $ Z[A] $, (\ref{3.3}) & \\
\hline
$\la$ & $\vp(x)$ & $<A_{op}>$, (\ref{3.4c}) &
$ \Zhat[\la] $, (\ref{3.4a}) & \\
:real & & $\labar(<A_{op}>)$, (\ref{3.4c}) & &
$\Gahat[\la]=\hbar~ln~\Zhat[\la] $, \\
& & $\Zhat[\labar(A)]$ & & (\ref{3.4g}) \\
\hline
$<A_{op}>$ & & & &
$ \Ga[<A_{op}>]$ , \\
& & $\labar(<A_{op}>)$,(\ref{3.4d}) & &
(\ref{3.4d}) \\
\hline
$A$ & $\vp(x)$, & $\vp_c(x,\la)$,(\ref{3.4i}) &
$Y[A,\la]$,(\ref{3.4f}) & \\
& $\la$ & $\la_c(A)=\labar(A)$ & & $\Ga^{eff}[A,\la]$
\\
&:complex & :real,(\ref{3.4l}) &
& $=ln~Y[A,\la]$,(\ref{3.4f}) \\
& & $\Zhat[\la_c(A)]$ & & \\
& & $\Ga^{eff}[A,\la_c]$ & & \\
& & $=\Ga[A]$,(\ref{3.4n}) & & \\
\hline
\multicolumn{5}{c}{\q} \\
\multicolumn{5}{c}{Table 1\q Some Thermodynamic Functions.}
\end{tabular}
\vspace{0.5cm}
$\qq$ We notice the area constraint ,which was expressed as the delta function
in (\ref{3.3}), is replaced by the $\la$-integral in the present
formalism using $Y[A,\la]$. This is the key point for
correctly treating the area constraint in the semiclassical method\cite{N2f}.
The quantum effect is systematically evaluated loop-wise\cite{S1}:\ the
renormalization of parameters involved in the theory, due to the quantum
interaction of the Weyl mode $\vp$, is done in
eq.(\ref{3.4j}).
$\qq$ In the following,
we will evaluate the leading order,i.e. order of $\hbar^0$.
\section{Classical Vacua of R$^2$-Gravity}
$\qq$ The classical configuration (solution) of
the special case
,$\be=0$, was well-known as the Liouville solutions.
(See \cite{SEI} for a recent review.) Furthermore,
in the context of 2 dim quantum gravity (or the string theory ) ,
the special case was
already examined by \cite{OV} and \cite{Z} .
We consider the general case:\ $\be$\ is
the arbitray real number with the dimension of (Length)$^2$.
$\qq$ The classical field equation (\ref{3.4i}) is given by
\begin{eqnarray}
\frac{\del S_\la[\vp]}{\del\vp}=\frac{1}{\ga}\pl^2\vp
+\be\{ e^{-\vp}(\pl^2\vp)^2-2\pl^2(e^{-\vp}\pl^2\vp)\}-\la e^\vp=0\pr
\label{3.6a}
\end{eqnarray}
We make the following assumption of constant-curvature
for the solution.
\begin{eqnarray}
-R|_{\vp_c}=\
e^{-\vp_c}\pl^2\vp_c
=\mbox{const}\equiv \frac{-\al}{A}\com \label{3.8}
\end{eqnarray}
where $\al$ is a dimensionless constant.
The above equation is the Liouville equation.
It is easy to find that the
solution (\ref{3.8}) satisfies (\ref{3.6a}) for
such real $\al$ that satisfies the following equation:
\begin{eqnarray}
\mbox{COND.1}\qqqq\al^2\be'-\frac{1}{\ga}\al-\la A=0\com\q
\be'\equiv\frac{\be}{A}\pr
\label{3.9a}
\end{eqnarray}
We may safely assume the spherical symmetry (in (x,y)-plane)
for a stable solution:
$\vp_c=\vp_c(r),r=\sqrt{x^2+y^2}$. Then (\ref{3.8}) reduces to
\begin{eqnarray}
\frac{1}{r}\frac{d}{dr}r\frac{d\vp_c}{dr}+\frac{\al}{A}e^{\vp_c}=0
\pr \label{3.7}
\end{eqnarray}
$\qq$Before further anlysis , we comment on the
eq.(\ref{3.8}). The Liouville equation
(\ref{3.8}) corresponds to the ordinary gravity ($\be=0$)
\ :\ $S=\intx (\frac{1}{2\ga}\vp\pl^2\vp-\mu_1~e^\vp)$\ ,
with the
cosmological constant $\mu_1=-\frac{1}{\ga}\frac{\al}{A}$,
which is negative for $\al>0$ and positive for $\al<0$.
\subsection{Positive Curvature Solution}
$\qq$ For the case:
\begin{eqnarray}
\al >0\com \label{3.10b}
\end{eqnarray}
the solution of (\ref{3.7}) is given by (cf.\cite{OV,Z,SEI}),
\begin{eqnarray}
\vp_c(r;\al )=-ln~\{ \frac{\al}{8}(1+\frac{r^2}{A})^2\}\pr \label{3.10a}
\end{eqnarray}
This solution satisfies
$\left.\intx\sqg\right|_{\vp_c}=\intx~e^{\vp_c}=\frac{8\pi}{\al}A$\ ,
$-\left.\intx\sqg R\right|_{\vp_c}=\intx~\pl^2\vp_c=-8\pi$\ .
It means the manifold described by (\ref{3.10a})
has the area $\frac{8\pi}{\al}A$\ and
is topologically the sphere.
The equations (\ref{3.9a}
-\ref{3.10a}) constitute a solution of (\ref{3.6a}).
$\qq$ $S_{\la}[\vp_c]$ is given as
\begin{eqnarray}
& S_{\la}[\vp_c]\left(=\Gahat^0(\la)\right)
=(1+\xi)\frac{4\pi}{\ga}~ln\frac{\al}{8}-16\pi\al\be'+C(A)\com & \nn\\
& C(A)=\frac{8\pi (2+\xi)}{\ga}+\frac{8\pi\xi}{\ga}
\{~ln(1+L^2/A)-(L^2/A)/(1+(L^2/A))~\} &
\label{3.11}\\
& =\frac{8\pi (2+\xi)}{\ga}+\frac{8\pi\xi}{\ga}
\{~ln~\frac{L^2}{A}-1~\}+O(\frac{A}{L^2})\com\q \frac{L^2}{A}\gg 1\com
&\nn
\end{eqnarray}
where
$L$\ is the infrared cut-off ($r^2\leq L^2$)
introduced for the divergent volume intgral of the total derivative term
($\xi$-term). See Fig.2.
The integral is log-divegent at $r \rightarrow\infty$\ for
the classical solution (\ref{3.10a}).
$\al$ (or $\la$)is rewritten in terms of $\la$ (or $\al$)
by use of (\ref{3.9a}) and $C(A)$\ does not depend on $\al$\ and $\be$\
but depends on $A$\ and $\ga$.
In the above derivation, formula in App.A are usefull.
{\vs 5}
\begin{center}
Fig.2\q Infra-red cut-off $L$\ in
the flat coordinates and the sphere manifold.
For simplicity, the picture is for $\al=8.$\ For general $\al$,
$(x,y,r,\sqrt{A},L)$\ is substituted by
$\sqrt{8/\al} \times (x,y,r,\sqrt{A},L)$.
\end{center}
$\qq$ The eq. (\ref{3.4l}) at the classical level is written as,
\begin{eqnarray}
\frac{dS_{\la}[\vp_c]}{d\la}+A
=(\frac{4\pi}{\ga}\frac{1}{\al}(1+\xi)
-16\pi\be')\frac{d\al}{d\la}+A \nn\\
=\{ \frac{4\pi}{\ga}\frac{1}{\al}(1+\xi)-(16\pi\be'+\frac{1}{\ga})
+2\be'\al\}\frac{d\al}{d\la}=0\com \label{3.16}
\end{eqnarray}
where we have used a relation :\ $1=\frac{d\la}{d\al}\frac{d\al}{d\la}
=\frac{1}{A}(~2\al\be'-\frac{1}{\ga})\frac{d\al}{d\la}\ $
, which is derived
from (\ref{3.9a}). For the case $\frac{d\al}{d\la}\neq 0$~\cite{N3a}
,(\ref{3.16}) says
\begin{eqnarray}
\mbox{COND.2} & \Xi(\al;\be',\ga,\xi)\equiv
2\be'\al^2-(16\pi\be'+\frac{1}{\ga})\al+
(1+\xi)\frac{4\pi}{\ga}=0\com & \nn\\
& \al^{\pm}_c=\frac{1}{4\be'}\{ 16\pi\be'+\frac{1}{\ga}
\pm\sqrt{D} ~\} \com & \label{3.17}\\
& D\equiv (16\pi\be')^2
+\frac{1}{\ga^2}-\xi\frac{32\pi\be'}{\ga}
=(16\pi\be'-\frac{\xi}{\ga})^2+\frac{1-\xi^2}{\ga^2}\pr & \nn
\end{eqnarray}
The relation (\ref{3.9a}) gives $\la^{\pm}_c(\be)
\equiv \la(\be,\al^{\pm}_c(\be))$. Note that the determinant of the above
quadratic equation ,$D$, is positive definite for all real
$\be$ if the following
condition is satisfied.
\begin{eqnarray}
-1\leq\ \xi\ \leq\ +1\pr \label{3.17b}
\end{eqnarray}
We consider this case in the following.
$\qq$ In summary, 2 unknown variables $\al$\ and $\la$\ ,are fixed by
two conditions COND.1 and 2 and they are expreesed by three physical
parameters $\be$\ ,$\ga$\ ,$A$\ and one free parameter $\xi$.
We list here the obtained result of important physical quantites.
\begin{eqnarray}
\mbox{Curvature} &
\al^{\pm}_c=\frac{4\pi}{w}\{ w+1~\pm\sqrt{w^2+1 -2\xi w} ~\}
\com\nn\\
\mbox{Classical Action} &
S_{\la}[\vp_c]
=(1+\xi)\frac{4\pi}{\ga}~ln~\frac{\al_c}{8}-\frac{w}{\ga}\al_c
+C(A) \com \nn\\
\mbox{String Tension} &
\la_c A=\frac{1}{16\pi\ga}({\al_c}^2w-16\pi\al_c)\com \label{3.18}\\
\mbox{An Expect. Value} &
-A<\intx\sqg R^2>|_{c}=
\frac{\pl \Ga^{eff}[\vp_c]}{\pl\be'} \nn\\
&\qq =-16\pi\al_c+\al_c^2+\Xi\times
\frac{1}{\al_c}\frac{d\al_c}{d\be'} \com \nn\\
\mbox{Free Energy} &
-\Ga^{eff}|_c=\ -S_{\la}[\vp_c]-\la_c A\com \nn
\end{eqnarray}
where $w\equiv 16\pi\be'\ga$\ and $C(A)$\ is given in (\ref{3.11})\cite{N3b}
{}.
Note that $\Xi=0$.
$\qq$ The curvature$\times A$\
(\ $=R(\vp_c)\times A=\al(w)$\ ),
the classical value of $A<\intx \sqg R^2>$\ (\ =
$-\frac{\pl \Ga^{eff}[\vp_c]}{\pl\be'}$)
, the string tension$\times \ga A\ =\ga\la_c A\ $ and
the total free energy $\times \ga$\ ($=-\ga\Ga^{eff}[A]
=-\ga (S_{\la}+\la_c A)$\ )
are plotted, for the $-$branch solution,
in Fig.3 ,4,5 and 6 respectively.
In the figures we take $\xi=0.99$~whose meaning will be explained in Sec.5,
and the curves for the negative curvature solution (see Sec.3.2) are
also plotted.
The asymptotic behaviours of these quantities will be listed in Sec.3.3.
We also plot the area
$\times \frac{1}{A}$\ (
$=\frac{1}{A}\intx e^{\vp_c}=\frac{8\pi}{\al}$\ ) as the function of $w$
in Fig.7.
It shows
,as far as the classical configuration is concerned,
the $\del$-function condition in (\ref{3.3})
is not satisfied except
for the $\al^+_c$-solution in the $\be$~(or $w$~) $\rightarrow +\infty$ region
where $\la\rightarrow +\infty$\ ,
and for the $\al^-_c$-solution in the $\be$~(or $w$~)
$\rightarrow -\infty$ region
where $\la\rightarrow -\infty$\ .
This has happened because we are approximating
the quantumly fluctuating manifold
by the simple classical sphere whose configuration is specified
only by the effective area
$\frac{1}{\al}$\ and the string tension $\la$\ .
This characteristically shows the present effective action approach using
$Y[A,\la]$\ (\ref{3.4f}).
{\vs 5}
\begin{center}
Fig.3\q $A\times$ Curvature $=\al(w)$\ ,\
Positive and Negative Curv. Sols., $-$Branch
\end{center}
{\vs 5}
\begin{center}
Fig.4a\q Log-Log Plot of $A<\intx\sqg R^2>|_c$, $w>0$,
Positive and Negative Curv. Sols., $-$Branch
\end{center}
{\vs 5}
\begin{center}
Fig.4b\q Linear Plot of $A<\intx\sqg R^2>|_c$,
Positive and Negative Curv. Sols., $-$Branch
\end{center}
{\vs 5}
\begin{center}
Fig.5\q$\ga A\times$(String Tension)$=\ga\la(w)A$,
Positive and Negative Curv. Sols., $-$Branch
\end{center}
{\vs 5}
\begin{center}
Fig.6\q$\ga \times$(Total Free Energy)$=-\ga\Ga^{eff}$,
Positive and Negative Curv. Sols., $-$Branch,
The $w$-independent terms ($C(A),\Ctil_2(A)$)
are omitted.
\end{center}
{\vs 5}
\begin{center}
Fig.7 \q $\frac{1}{A}\times$Area$=\frac{8\pi}{\al(w)}$,
Positive and Negative Curv. Sols., $-$Branch
\end{center}
\subsection{Negative Curvature Solution}
$\qq$ For the case:
\begin{eqnarray}
\al <0\com \label{4.1}
\end{eqnarray}
the solution of (\ref{3.8}) is given by (cf.\cite{SEI}),
\begin{eqnarray}
\vp_n(r;\al )=-ln~\{ \frac{-\al}{8}(1-\frac{r^2}{A})^2\}\pr \label{4.2}
\end{eqnarray}
The equations (\ref{3.9a}-\ref{3.7}) and
(\ref{4.1}-\ref{4.2}) constitute another solution of (\ref{3.6a}).
It is singular at $r=\sqrt{A}$~, which means
the manifold is open. There exist two independent regions:\ the inner region
\ $0\leq r<\sqrt{A}$\ and the outer region\ $r>\sqrt{A}$. We consider only the
inner region\cite{N3c}
{}.
We should carefully treat the singularity
by introducing a proper regularization. We regularize the inner region
by $0\leq r<\sqrt{A-\ep}$~ where $\ep\rightarrow +0$~. See Fig.8.
{\vs 5}
\begin{center}
Fig.8\ Reguralization of singularity at $r=\sqrt{A}$ \ of (\ref{4.2}).
\end{center}
Then various terms in the action are evaluated as
\begin{eqnarray}
\intx~e^{\vp_n}
=\lim_{\ep\rightarrow +0}\int_{0\leq r^2 <A-\ep}
{}~d^2x~e^{\vp_n}
=\frac{8\pi}{-\al}A\lim_{\ep\rightarrow +0}
(\frac{A}{\ep}-1)\com \nn\\
\intx~\vp_n\pl^2\vp_n
=8\pi \lim_{\ep\rightarrow +0}\{2-(\frac{A}{\ep}-1)ln~(\frac{-\al}{8})
+\frac{2A}{\ep}(ln\frac{A}{\ep}-1)\ \}\com \label{4.3}\\
\intx~e^{-\vp_n}(\pl^2\vp_n)^2
=\frac{-8\pi\al}{A}\lim_{\ep\rightarrow +0}
(\frac{A}{\ep}-1)\com \nn\\
-\left.\intx\sqg R\right|_{\vp_n}=\intx~\pl^2\vp_n
=8\pi\lim_{\ep\rightarrow +0}(\frac{A}{\ep}-1)\com \nn\\
\intx\pl_a\vp_n\pl_a\vp_n=16\pi
\lim_{\ep\rightarrow +0}(~ln\frac{\ep}{A}+\frac{A}{\ep}-1)\com\nn
\end{eqnarray}
where $\ep$ \ is introduced as a regularization parameter and is
a positive infinitesimally-small constant with the dimension of area
\cite{N3cc}.
The divergence of the total area $\intx~e^{\vp_n}$
, at this stage, says the manifold
considered is not closed. The singular point $r=\sqrt{A}$\ corresponds
to the boundary of the open manifold.
$\qq$ Using the above results, the Euclidean action (\ref{3.4g}) and
eq. (\ref{3.4l}), at the classical level, are given by
\begin{eqnarray}
S_{\la}[\vp_n]=-\frac{4\pi}{\ga}(1+\xi)\La~ ln~(\frac{-\al}{8})
+8\pi\al\be'\La
+\frac{8\pi}{\al}\la A \La +C_2(A) \com \nn\\
C_2(A)=\frac{4\pi(1+\xi)}{\ga}\{ 2+2(\La+1)(ln(\La+1)-1)\}
+\frac{8\pi\xi}{\ga}(-ln(\La+1)+\La)\com\nn\\
\La\equiv \lim_{\ep\rightarrow +0}(\frac{A}{\ep}-1)>0\com \nn\\
\frac{dS_{\la} [\vp_n]}{d\la}+A=
\{ -\frac{4\pi\La}{\ga\al}(1+\xi)+8\pi\be'\La-\frac{8\pi}{\al^2}\la A\La\}
\frac{d\al}{d\la}+\frac{8\pi}{\al}A\La+A \label{4.4}\\
=\{-\frac{4\pi\La}{\ga\al}(1+\xi)+8\pi\be'\La-\frac{8\pi}{\al^2}\la A\La
+(2\al\be'-\frac{1}{\ga})(\frac{8\pi\La}{\al}+1)\}\frac{d\al}{d\la}=0
\pr\nn
\end{eqnarray}
The above equations are divergent and are not well-defined
for $\ep\rightarrow +0$.
We can,however,absorb the divergence by the rescaling of the
coupling $\be'$\ ,the curvature parameter $\al$\ ,the cosmological
parameter $\la$\ and some physical operators to be evaluated.
\begin{eqnarray}
\altil\equiv \frac{\al}{\La}\com\q \betil'\equiv\La\be'\ (\betil\equiv\La\be)
\com \label{4.5}
\end{eqnarray}
where $\La$~is the positive divergent constant introduced in (\ref{4.4}).
Note that $\al$ \ does not change its sign by this transformation:\
$\altil<0$\ .The corresponding transformation of $\la$\ is obtained by
the requirement of keeping the form of (\ref{3.9a}).
\begin{eqnarray}
\mbox{COND.}{\tilde 1}\qqqq
\latil A\equiv \frac{\la}{\La}A=\altil^2\betil'-\frac{\altil}{\ga}\pr
\label{4.6}
\end{eqnarray}
$\qq$ In terms of rescaled quantities, we can rewrite (\ref{4.4}) as
\begin{eqnarray}
\Stil_{\latil}[\vp_n]\equiv\frac{S_{\la}[\vp_n]}{\La}
=-\frac{4\pi}{\ga}(1+\xi)~ln~(\frac{-\altil}{8})+16\pi\altil\betil'
+\Ctil_2(A) \com \nn\\
\Ctil_2(A)=\frac{8\pi(1+\xi)}{\ga}ln~\La-\frac{8\pi}{\ga} \com \label{4.7} \\
\frac{dS_{\la} [\vp_n]}{d\la}+A=
\frac{d\Stil_{\latil}[\vp_n]}{d\latil}+A=
\frac{d\altil}{d\latil}[-\frac{4\pi}{\ga}\frac{1}{\altil}(1+\xi)+16\pi\betil'
+2\altil\betil'-\frac{1}{\ga}]=0\ .\nn
\end{eqnarray}
This result shows the rescaled action is finite except a constant term
(log-divergent) and the equation for $\latil_n$\ or $\altil_n$
\ (i.e.\ eq.(\ref{3.4l}))
, at the classical level, is completely
free from divergence.
\begin{eqnarray}
\mbox{COND.}{\tilde 2} &{\tilde \Xi}(\altil;\betil',\ga,\xi)\equiv
2\betil'\altil^2+(16\pi\betil'-\frac{1}{\ga})\altil
-\frac{4\pi}{\ga}(1+\xi)=0\com \nn\\
& \altil^{\pm}_n=\frac{1}{4\betil'}\{ -16\pi\betil'+\frac{1}{\ga}
\pm\sqrt{D } ~\}
\com \label{4.8}\\
&
D\equiv (16\pi\betil')^2+\frac{1}{\ga^2}+32\pi\frac{\betil'}{\ga}\xi
=(16\pi\betil'+\frac{\xi}{\ga})^2+\frac{1}{\ga^2}(1-\xi^2)\pr \nn
\end{eqnarray}
COND.${\tilde 1}$\ (\ref{4.6})\ gives
$\latil^\pm_n(\betil)=\latil(\betil,\altil^\pm_n(\betil))$.
We consider,again, the following region of $\xi$\ ,in order to
guarantee $D\geq 0$~ for all real $\betil$.
\begin{eqnarray}
-1\leq\xi\leq 1\q\pr \label{4.8b}
\end{eqnarray}
The physical quantity of $<\intx~\sqg R^2>|_n$\ is given by
\begin{eqnarray}
-\frac{A}{\La^2}<\intx\sqg R^2>|_{n}=
\frac{1}{\La^2}\frac{d\Ga^{eff}[\vp_n]}{d\be'}|_n
=\frac{d{\tilde \Ga}^{eff}[\vp_n]}{d\betil'}|_n \nn\\
=+16\pi\altil_n+\altil_n^2+
{\tilde \Xi}\times\frac{1}{\altil_n}\frac{d\altil_n}{d\betil}\com\q
\label{4.9}
\end{eqnarray}
where ${\tilde \Xi}=0$.
$\qq$ We note the difference in signs between the equations in the positive-
curvature case, (\ref{3.17}-\ref{3.18}), and those in the negative-curvature
case, (\ref{4.6}-\ref{4.9}).
Remarkably,by the following sign change,
\begin{eqnarray}
\betil\rightarrow -\betil\com
\altil\rightarrow -\altil\com
(\q\latil\rightarrow -\latil\com
\Stil_{\latil}-\Ctil_2(A)\rightarrow -(\Stil_{\latil}-\Ctil_2(A))
\com\q) \label{4.10}
\end{eqnarray}
the above negative-curvature results
($\altil_n,\Stil_{\latil}-\Ctil_2(A),\latil_n$)
as the functions of $\betil'$\ ,
reduce to the same forms of the positive-curvature ones
($\al_c,S_{\la}-C(A),\la_c$) as the functions of $\be'$\cite{N3d}
{}.
$\qq$ We show the behaviours of $\altil^{\pm}_n,
-\frac{\pl{\tilde \Ga}^{eff}_{\pm}}{\pl\betil'}$\ , $\ga\latil^{\pm}_n A$\
and $-\ga{\tilde \Ga}^{eff}_\pm (\ =-\ga(\Stil^\pm_\latil +\latil^\pm_n A)\ )$
\
in the dotted lines of
Fig.3,4,5 and 6 respectively. The figures show the above reflection
symmetry clearly.
The asymptotic behaviours of the above physical quantities will be
listed in Table 3 of Sec.3.3.
It is very interesting that we can define finite quantities
in the open manifold in the above rescaling procedure
(\ref{4.5}-\ref{4.7}). It reminds us of the renormalization in the
quantum field theory. The present case is, however, a procedure to absorb
the infrared divergence due to the coordinate singularity of the classical
open manifold, not to absorb the ultraviolet divergence in the quantum theory.
Note that the constant-curvature sign remains negative after the rescaling
and the Euler number $\intx\sqg R$~ is negatively divergent.
These facts make us envisage Fig.9 as the rescaled manifold.
It describes a sphere punctured over the surface. Each puncture absorbs
the infrared divergence.
{\vs 5}
\begin{center}
Fig.9\ Punctured sphere absorbing infrared divergence
\end{center}
\subsection{ Phases and Asymptotic Behaviours}
$\qq$ The asymptotic behaviours of the physical quantites,obtained in Sec.3.1,
are listed in Table 2, where the case of $\al<0$~
is excluded due to the present condition (\ref{3.10b}).
\vspace{0.5cm}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
& & $w\ll -1$ & $-1\ll w<0$
& $0<w\ll 1$
& $1\ll w$ \\
\hline
& Phase & /&/&(E)&(D) \\
\cline{2-6}
& $\al^+_c$ &$<0$\ ,
& $<0$\ ,
& $\frac{4\pi}{w}\{2+w(1-\xi)$
& $4\pi\{2+\frac{1-\xi}{w} $ \\
& &\ not allowed & \ not allowed
& $+O(w^2)\}$ & $+O(w^{-2})\}$ \\
\cline{2-6}
$+$ & $-\frac{\pl \Ga^{eff}_+}{\pl\be'}$ &/
& / & $-\frac{64\pi^2}{w^2}\{ 1-(1+\xi)w$
& $64\pi^2\{1+\frac{0}{w}$ \\
& & & &$\ \ +O(w^2)\}$ & $+O(w^{-2})\}$ \\
\cline{2-6}
& $\ga\la^+_cA$&/
& / & $\frac{4\pi}{w}\{-1+0\cdot w$
& $4\pi w\{1$ \\
& & & & $+O(w^2)\}$ & $+O(w^{-1})\}$ \\
\cline{2-6}
& & /&/& $\frac{4\pi}{w}\{1+2w$ & $4\pi w\{1 $ \\
& $-\ga \Ga^{eff}_+$ & & & $+(1+\xi)w~ln~w$ & $+O(w^{-1})\} $ \\
& & & & $+O(w^2)\}-\ga C(A)$ & $-\ga C(A) $ \\
\hline
& Phase & (C) & \multicolumn{2}{c|}{(B)} & (A) \\
\cline{2-6}
& $\al^-_c$ & $8\pi$
& \multicolumn{2}{c|}{ $4\pi(1+\xi)\{1$ }
& $\frac{4\pi(1+\xi)}{w}$ \\
& & $+O(|w|^{-1})\}$ & \multicolumn{2}{c|}{$-\frac{1-\xi}{2}w\}+O(w^2)$}
& $+O(w^{-2})$ \\
\cline{2-6}
$-$ & $-\frac{\pl \Ga^{eff}_-}{\pl\be'}$ & $64\pi^2+\frac{0}{|w|}$
& \multicolumn{2}{c|}{
$16\pi^2(1+\xi)\{3-\xi$ }
& $\frac{64\pi^2(1+\xi)}{w} $ \\
& & $+O(w^{-2})$
& \multicolumn{2}{c|}{
$-(1-\xi)^2w\}+O(w^2)$ }
& $+O(w^{-2})$ \\
\cline{2-6}
& $\ga\la^-_cA$ &$-4\pi |w|\{1$
& \multicolumn{2}{c|}{ $4\pi(1+\xi)\{-1$ }
& $-\frac{\pi}{w}(1+\xi)(3-\xi)$ \\
& & $+O(\frac{1}{|w|})\}$
& \multicolumn{2}{c|}{$+\frac{3-\xi}{4}w\}+O(w^2)$}
& $+O(w^{-2}) $ \\
\cline{2-6}
& & $-4\pi |w|\{1$ & \multicolumn{2}{c|}{
$4\pi(1+\xi)\{1-ln~\frac{1+\xi}{2} $
}
& $4\pi(1+\xi)~ln~w$ \\
& $-\ga\Ga^{eff}_-$ & $+O(\frac{1}{|x|})\} $
& \multicolumn{2}{c|}{ $+\frac{3-\xi}{4}w\}+O(w^2)$ }
& +const \\
& & $-\ga C(A)$ & \multicolumn{2}{c|}{$-\ga C(A)$} & $-\ga C(A) $ \\
\hline
\multicolumn{6}{c}{\q} \\
\multicolumn{6}{c}{Table 2\ \ Asymp. behaviour of physical quantities.}\\
\multicolumn{6}{c}{
$R>0, w\equiv 16\pi\be'\ga, \ga=\frac{48\pi}{26-c_m}>0\ (c_m<26)$.
$C(A)$ is given by (\ref{3.11}).}
\end{tabular}
\vspace{0.5cm}
$\qq$ Due to the 'reflection symmetry',
each phase of the negative-curvature solution,
given in Sec.3.2, is characterzed
in the similar way as in the positive-curvature case.
We list the phase characterization in Table 3.
('Primes' in Table 3 mean modification due to the sign difference.)
\begin{tabular}{|c|c|c|c|c|}
\hline
& $w\ll -1$ & $-1\ll w<0$
& $0<w\ll 1$
& $1\ll w$ \\
\hline
+ & (D') & (E') & / & / \\
\hline
- & (A') & \multicolumn{2}{c|}{(B')} & (C') \\
\hline
\multicolumn{5}{c}{\q} \\
\multicolumn{5}{c}{Table 3\q Phases of Negative Curvature Solution.
$w\equiv 16\pi\betil'\ga$.}
\end{tabular}
\vspace{0.5cm}
$\qq$ All phases are explained in \cite{ITY1}
using the above asymptotic behaviour.
In the present paper we will characterize each phase by the field equation
satisfied in each asymptotic region in Sec.4 and by
the equation of state in Sec.7.
\vspace{1cm}
\section{Asymptotic Regions}
$\qqq$Now we consider the classical solutions of (\ref{3.6a}) in the
asymptotic regions\ :\
(a) $|w|\rightarrow \infty$\ ;\
(b) $|w|\rightarrow +0$\ .
Table 2 and 3 in Sec.3.3 say each region has two cases.
\flushleft{(ai)\ $|\ga\la A|\sim O(\frac{1}{w})$\ ,\
Weak-Field Vacua, ((A),(A'))}
In this region, the following parts of (\ref{3.6a}) are dominant.
\begin{eqnarray}
e^{-\vp}(\pl^2\vp)^2-2\pl^2(e^{-\vp}\pl^2\vp)=0\pr\nn\\
\vp_{ary}=\mbox{const}\com\ R(\vp_{asy})=0\pr
\label{4.3.5}
\end{eqnarray}
These vacua are defined only by the 'kinetic terms' in the action.
Therefore we call these vacua ((A),(A'))
{\it Weak-Field}(WF-){\it Vacua}\cite{N4a}.
\flushleft{(aii)\ $\ga\la A\sim O(w)$\ ,\ Perfect Sphere Vacua,
((C),(C'),(D),(D'))}
\begin{eqnarray}
e^{-\vp}(\pl^2\vp)^2-2\pl^2(e^{-\vp}\pl^2\vp)-c~e^\vp=0\pr\nn\\
-R|_{asy}=e^{-\vp_{asy}}\pl^2\vp_{asy}=\mbox{const}=\pm\sqrt{c}(\neq 0)\pr
\label{4.3.6}
\end{eqnarray}
For the positive curvature case (lower sign case), $c$\ can be fixed
,by the condition $\intx\sqg R=8\pi$\ ,as $\sqrt{c}=\frac{8\pi}{A}$\
and the total area is $\frac{8\pi}{R}=A$\ .
These vacua
are strongly restricted by the 'potential term' of $e^\vp$\ and
describe a perfect sphere.
We call the vacua (D) and (C') {\it expansive perfect sphere}
where the string tension positively divergent, and
the vacua (C) and (D') {\it tensed perfect sphere}
where the string tension negatively divergent.
\flushleft{(bi)\ $\ga\la A\sim \mbox{const}$\ ,
Liouville Vacua, ((B),(B'))}
\begin{eqnarray}
\frac{1}{\ga}\pl^2\vp-\la(0)~e^\vp =0\com\q\mbox{Liouville Eq.}\q\com\nn\\
R=-e^{-\vp}\pl^2\vp=-\ga\la(0)\pr
\label{5b.1}
\end{eqnarray}
This corresponds to the $\be=0$~theory.
In Phase (B), the Euler number is properly given by
$\intx\sqg R=-\ga\la(0)\cdot\intx\sqg=-\ga\la(0)A\cdot\frac{8\pi}{\al(0)}
=4\pi(1+\xi)\cdot\frac{8\pi}{4\pi(1+\xi)}=8\pi$~
(for arbitrary $\xi$~).
We call these regions Liouville Vacua.
\flushleft{(bii)\ $\ga\la A\sim O(\frac{1}{w})$\ ,
Degenerate Vacua, ((E),(E'))}
In these regions the curvature must depend on $w$~ in order that
Eq.(\ref{3.6a}) is satisfied.
\begin{eqnarray}
\frac{1}{\ga}e^{-\vp}\pl^2\vp
+\be\{ (e^{-\vp}\pl^2\vp)^2-2e^{-\vp}\pl^2(e^{-\vp}\pl^2\vp)\}-\la(\be) =0\com
\nn\\
R=-e^{-\vp}\pl^2\vp\sim\ \frac{1}{A}\times O(\frac{1}{w})\pr\label{5b.2}
\end{eqnarray}
All terms of (\ref{3.6a}) are effective.
Because the total area $\frac{8\pi}{R}$~ vanishes, we name these regions
{\it degenerate vacua}.
\vspace{0.5cm}
We list all above asymptotic regions in Table 3 with the effective terms
of (\ref{3.6a}) marked by $\bigcirc$.
\begin{tabular}{|c|c|c|c|c|}
\hline
\ \ Terms of (\ref{3.6a}) & $\frac{1}{\ga}\pl^2\vp$ &
$+\be\{ e^{-\vp}(\pl^2\vp)^2-2\pl^2(e^{-\vp}\pl^2\vp)\}$ &
$-\la e^\vp $ & \\
\hline
Weak-Field & - & $\bigcirc$ & - & (ai) \\
\hline
Perf.Sphere & - & $\bigcirc$ & $\bigcirc$ & (aii) \\
\hline
Liouville & $\bigcirc$ & - & $\bigcirc$ & (bi) \\
\hline
Degenerate & $\bigcirc$ & $\bigcirc$ & $\bigcirc$ & (bii) \\
\hline
Free Boson & $\bigcirc$ & - & - & non-exist \\
\hline
& $\bigcirc$ & $\bigcirc$ & - & non-exist \\
\hline
\multicolumn{5}{c}{\q} \\
\multicolumn{5}{c}{Table 3\q Asymptotic States }
\end{tabular}
\vspace{1cm}
Now we have characterized all asymptotic regions. We can see,
as shown in Fig.3,
$\al^-_c$ solution connects between WF-vacuum at $w=+\infty$\
and the tensed perfect sphere vacuum at $w=-\infty$. And
$\al^-_n$ solution connects between WF-vacuum at $w=-\infty$\
and the expansive perfect sphere vacuum at $\be=+\infty$.
\section{ $\la$-integral }
In this section we do the $\la$-integral of (\ref{3.4k}) in the
lowest order. This part gives us some contribution to $Z[A]$.
We consider the positive curvature solution.
After splitting $\la$~ around $\la_c$~ :\ $\la=\la_c+\om$~
, the $\om$-integral part of
(\ref{3.4ll}) is approximated as
\begin{eqnarray}
& Z_{\om}[A]
\equiv\int d\om~exp~\{~\frac{1}{2}
\left.\frac{d^2\Gahat[\la]}{d\la^2}\right|_{\la_c}~\om^2
+O(\om^3) \ \} & \nn\\
& \approx
\int d\om~exp~\{~\frac{1}{2}
\left.\frac{d^2S_\la}{d\la^2}\right|_{\la_c}~\om^2\ \}\pr &\label{la.1}
\end{eqnarray}
{}From (\ref{3.11}) and (\ref{3.9a}), we can obtain
\begin{eqnarray}
& \frac{d^2S_\la[\vp_c]}{d\la^2}=
-\frac{4\pi}{\ga}\frac{1+\xi}{\al^2}(\frac{d\al}{d\la})^2
+(\frac{4\pi}{\ga}\frac{1+\xi}{\al}
-16\pi\be')\frac{d^2\al}{d\la^2} & \nn\\
& =A^2\frac{4\pi}{\al^2}\frac{1}{(2\al\be'-\frac{1}{\ga})^3}
\{~4(1-\xi)\al^2\be'^2+(1+\xi)(2\al\be'-\frac{1}{\ga})^2~\}
& \label{la.2}\\
& =A^2\frac{4\pi}{\al^2}\frac{\ga}{(\frac{\al w}{8\pi}-1)^3}
\{~(1-\xi)(\frac{\al w}{8\pi})^2+(1+\xi)(\frac{\al w}{8\pi}-1)^2~\}\com &\nn
\end{eqnarray}
where we have used some relations derived from (\ref{3.9a})~:~
$d\al/d\la=A/(2\al\be'-\frac{1}{\ga})\com\
d^2\al/d\la^2=-2\be'A^2/(2\al\be'-\frac{1}{\ga})^3\pr$\
Putting $\al^\pm_c$-solution of (\ref{3.18}) into the above expression,
we can confirm
$d^2S_\la/d\la^2|_{\al^-_c}< 0\ ,\
d^2S_\la/d\la^2|_{\al^+_c}> 0 $~ for all $\be$~(or $w$) region.
(Note that $\{\ \}$-part of (\ref{la.2}) is positive definite.)
For the $\al^-_c$-solution, it is necessary to
change the integeral path from the original pure imaginary $\om$~ in Fig.1
to the real $\om$~ as shown in Fig.10. $Z_\om[A]$~ is evaluated as
\begin{eqnarray}
& Z_\om[A]=\frac{1}{\sqrt{-\frac{d^2S_\la}{d\la^2}|_{\al^-_c}} }
\q \mbox{for $-$branch solution} & \nn\\
& Z_\om[A]=\frac{1}{\sqrt{+\frac{d^2S_\la}{d\la^2}|_{\al^+_c}} }
\q \mbox{for $+$branch solution} & \pr \label{la.3}
\end{eqnarray}
{\vs 7}
\begin{center}
Fig.10\ $\la$-integral path for $-$branch solution
\end{center}
Using the results of Table 2, the asymptotic behaviours are evaluated as
\begin{eqnarray}
\mbox{Phase (A)}\q w\gg 1\com &
\ln~Z_\om\sim -\ln~A-\ln~w\com \nn\\
\mbox{Phase (B)}\q |w|\ll 1\com &
\ln~Z_\om\sim -\ln~A+\mbox{const}\com \nn\\
\mbox{Phase (C)}\q w\ll -1\com &
\ln~Z_\om\sim -\ln~A+\half\ln~|w|\com \label{la.4}\\
\mbox{Phase (D)}\q w\gg 1\com &
\ln~Z_\om\sim -\ln~A+\half\ln~w\com \nn\\
\mbox{Phase (E)}\q 0<w\ll 1\com &
\ln~Z_\om\sim -\ln~A+\half\ln~w\pr \nn
\end{eqnarray}
We notice the first term of each right-hand side contributes to the
string susceptibility (see Sec.6). The second term does not
have a factor of $4\pi/\ga=(26-c_m)/12$~ in comparison with the
$\Ga^{eff}$~ of Table 2. Because the factor means the number of
freedom in this thermodynamical system (see Sec.7) , their contribution
is negligible except for the case :\ $c_m\approx 26$.
\vspace{1cm}
\section{ Cross-Over Points and Determination of $\xi$}
$\qq$ Let us see the $\xi$-dependence of the cross-over points.
Because the negative constant curvature solution is obtained
by the reflection symmetry from the positive one, we discuss
only the latter one.
As for the +branch solution, the string tension
$\la$\ changes
its sign at $w_0$\ ,which is located somewhere between
(D)-phase and (E)-phase of Table 2(see Fig.5)\ .
We can obtain it as the zero of
$\la^+_c(w_0)=0$\ in (\ref{3.18}).
\begin{eqnarray}
w_0(\xi)=\frac{4}{3-\xi}\pr \label{3.19}
\end{eqnarray}
The - solution has two cross-over points.
The log-log plot of
$-\frac{\pl \Ga^{eff}_-[\vp_c]}{\pl\be'}$ (Fig.4a) shows, at some point
$w_c>0$\ between phase (A) and (B), the behaviour changes from the
linearly-descending line to
the constant-line as we decrease $w$.
The linear plot of
$-\frac{\pl \Ga^{eff}_-[\vp_c]}{\pl\be'}$ (Fig.4b) shows, at some point
$w_c'<0$\ between phase (B) and (C), the behaviour changes from the
linearly-descending line to
the constant-line as we decrease $w$ in the negative region.
Those straight lines can be obtained as
\begin{eqnarray}
-\frac{\pl \Ga^{eff}_-[\vp_c]}{\pl\be'}
\rightarrow 64\pi^2\frac{1+\xi}{w}\q\mbox{as}\q w\rightarrow +\infty
\com \nn\\
-\frac{\pl \Ga^{eff}_-[\vp_c]}{\pl\be'}
\rightarrow 16\pi^2(1+\xi)\{(3-\xi)-(1-\xi)^2w
+O(w^2)\}\q\mbox{as}\q w\rightarrow +0\com \label{3.20} \\
-\frac{\pl \Ga^{eff}_-[\vp_c]}{\pl\be'}
\rightarrow 64\pi^2+\frac{0}{|w|}+O(w^{-2})\q\mbox{as}\q w\rightarrow -\infty
\pr \nn
\end{eqnarray}
We can clearly define the changing points $w_c$\ and $w_c'$\
as the cross-point of two corresponding
asymptotic lines above, and obtain as
\begin{eqnarray}
w_c(\xi)=\frac{4}{3-\xi}= w_0(\xi)\com\q
w_c'(\xi)=-\frac{1}{1+\xi}\pr \label{3.21}
\end{eqnarray}
All cross-over points depend on
the parameter of the total derivative
term ,$\xi$\ ,and which says the global term controls the essential
behaviour of the theory.
$\qq$ What value should we take for $\xi$~? It can be answered
, purely within the theory, from the quantum analysis\cite{S1}.
When we take $\xi=1$~, the renormalization-group beta functions
have zeros for $w\geq 1$~. Here, however,
we fix the parameter $\xi$~ by adjusting the asymptotic
(\ $A\rightarrow\infty $\ )
behaviour of $Z[A]$, for the case $\be=0$\ ,
with the KPZ (conformal) result\cite{KPZ}. The asymptotic behaviour of $Z[A]$
for $\be=0$\ is given as
\begin{eqnarray}
\al^-_c-\mbox{solution} \nn\\
Z[A]|_{w=0}\sim
A^{-\frac{8\pi\xi}{\ga}}\cdot A^{-1}=A^{-\frac{26-c_m}{6}\xi-1}\com\nn\\
\mbox{as}\ A\rightarrow +\infty\com
\label{4.3.1}
\end{eqnarray}
where the factor $A^{-1}$~ comes from $Z_\om$~ in Sec.5.
The KPZ result\cite{KPZ} is
\begin{eqnarray}
Z^{KPZ}[A]\sim A^{\ga_s-3}\com\
\ga_s=\frac{1}{12}\{c_m-25-\sqrt{(25-c_m)(1-c_m)}\}+2\pr \label{4.3.2}
\end{eqnarray}
In order to adjust our result with the KPZ result in the classical
limit $c_m\rightarrow -\infty$\ :\
$Z^{KPZ}[A]\sim A^{+\frac{1}{6}c_m}$\ , we must take
\begin{eqnarray}
\xi=1\pr \label{4.3.3}
\end{eqnarray}
Taking $\xi=1$, the asymptotic behaviour of $Z[A]$\ for the
$\al^-_c$-solution is
\begin{eqnarray}
Z[A]\sim A^{-\frac{26-c_m}{6}-1}\com\q A\rightarrow +\infty
\pr \label{4.3.4}
\end{eqnarray}
Now we compare the KPZ result and the semiclassical result in the
normalized form.
\begin{eqnarray}
Z^{KPZ}_{norm}[A]\equiv \frac{Z^{KPZ}[A]}{Z^{KPZ}[A]|_{c_m=0}}
\sim A^{\ga_s(c_m)-\ga_s(c_m=0)}\com\nn\\
\ga_s(c_m)-\ga_s(c_m=0)=\frac{1}{12}\{c_m+5-\sqrt{(25-c_m)(1-c_m)}\}
\com \label{4.3.4b}\\
Z_{norm}[A]\equiv \frac{Z[A]}{Z[A]|_{c_m=0}}
\sim A^{+\frac{c_m}{6}}\com\q \pr\nn
\end{eqnarray}
In Fig.11, the present semiclassical result and the KPZ result
are plotted.
{\vs 5}
\begin{center}
Fig.11\ Semiclassical result versus KPZ formula for string susceptibility
\end{center}
In the following, we take $\xi=1$\cite{N6a}
{}.
\vspace{1cm}
\section{ Phases, Thermodynamic Properties
and Equation of State}
$\qq$ In this section we examine the thermodynamic properties of the system
using the obtained analytic expression.
We consider the positive curvature solution.
The partition function is given by
\begin{eqnarray}
Z[A]=\int d\la~exp~\{~\Gahat[\la]+\la A~\}
\approx exp~\{\Gahat[\la_c]+\la_c A\}\com\nn\\
\frac{d}{d\la}(\Gahat[\la]+\la A)|_{\la_c}
=\frac{d\Gahat[\la_c]}{d\la_c}+A=0\pr \label{state.1}
\end{eqnarray}
Under the variation of the total area:\ $A\ra A+\Del A$~,
$\ln~Z[A]$~ changes by
$\Del (\ln~Z[A])=\la_c\cdot\Del A+\Del A\cdot\frac{d\la_c}{dA}\cdot
(\frac{d\Gahat}{d\la}+A)|_{\la_c}=\la_c\cdot\Del A$~.
Because the free energy $F$~ is given by
$F=- ln~Z[A]$~, the pressure $P$~ is obtained as
\begin{eqnarray}
& P=-\frac{\pl}{\pl A}F=\frac{\pl}{\pl A}\ln~Z[A]=\la_c
\pr & \label{state.2}
\end{eqnarray}
The pressure is the same as the string tension.
$\qq$ We define the temerature $T(w)$~ , imitating the Boyle-Charles' law,
as follows.
\begin{eqnarray}
& P\cdot A\equiv \frac{4\pi}{\ga}~T(w) \com & \nn\\
& T(w)=\frac{\ga\la_c A}{4\pi}=\frac{1}{64\pi^2}({\al_c}^2 w-16\pi\al_c)
\com & \label{state.3}\\
& \al_c(w)=\left\{ \begin{array}{ll}
\frac{4\pi}{w}\{ w+1~+|w-1| ~\} & \mbox{for + branch solution}\\
\frac{4\pi}{w}\{ w+1~-|w-1| ~\} & \mbox{for $-$branch solution}
\end{array}
\right. & \nn
\end{eqnarray}
$N\equiv 4\pi/\ga=(26-c_m)/12$~ corresponds to the 'mol number'.
The temperature is the (dimensionless) string tension per a unit mol.
The final analytic form of the temperature is given by
\begin{eqnarray}
+\mbox{branch solution}\qqq
& T(w)=\left\{ \begin{array}{ll}
-\frac{1}{w} & \mbox{for}\ 0<w\leq 1\\
w-2 & \mbox{for}\ 1\leq w
\end{array}
\right. & \nn\\
-\mbox{branch solution}\qqq
& T(w)=\left\{ \begin{array}{ll}
w-2 & \mbox{for}\ w\leq 1\\
-\frac{1}{w} & \mbox{for}\ 1\leq w
\end{array}
\right. & \label{state.4}
\end{eqnarray}
The behaviour of $T(w)$~ is plotted in Fig.12.
{\vs 5}
\begin{center}
Fig.12\ Temperature $T=T(w)$, Pos. Curv. Sol.
\end{center}
$\qq$ The asymptotic form of temperature in each phase is given by,
\begin{eqnarray}
\mbox{Phase (A)}\q w\gg 1\com &
P\cdot A=-\frac{4\pi}{\ga}\frac{1}{w}\com &
T_{(A)}= -\frac{1}{w}\com \nn\\
\mbox{Phase (B)}\q |w|\ll 1\com &
P\cdot A=-\frac{8\pi}{\ga}(1+O(w))\com &
T_{(B)}\approx -2\com \nn\\
\mbox{Phase (C)}\q w\ll -1\com &
P\cdot A=\frac{4\pi}{\ga}w(1+O(w^{-1}))\com &
T_{(C)}\approx w\com \label{state.5}\\
\mbox{Phase (D)}\q w\gg 1\com &
P\cdot A=\frac{4\pi}{\ga}w(1+O(w^{-1}))\com &
T_{(D)}\approx w\com \nn\\
\mbox{Phase (E)}\q 0<w\ll 1\com &
P\cdot A=-\frac{4\pi}{\ga}\frac{1}{w}\com &
T_{(E)}= -\frac{1}{w}\pr \nn
\end{eqnarray}
The negativeness of temperature, in Phase (A),(B) and (C),
says the matter-gass particles
attract each other.
The small absolute value of $T_{(A)}$~ says the matter-gass particles
move almost freely.
We can do the same analysis for the negative curvature solution.
The corresponding temperture is obtained by reflecting the graph
of Fig.12 following (\ref{4.10}).
It is interesting that
the matter-gass particles atracting each other on an open manifold
can be regarded as the 'repulsive'
particles on a regularized closed manifold.
$\qq$ The entropy is similarly obtained. Using the relation:\
$\Del w|_{A:fixed}=w\cdot \frac{\Del\be}{\be}\com\q
\Del T|_{A:fixed}=\frac{\pl T}{\pl w}\cdot w\frac{\Del\be}{\be}$\ ,
it is given as
\begin{eqnarray}
& S_{ent}=-(\frac{\pl F}{\pl T})_A &\nn\\
&= +\frac{1}{w}\frac{\pl w}{\pl T}\cdot \be\frac{\pl}{\pl\be}\ln~Z[A]
=-\frac{1}{16\pi\ga}\frac{\pl w}{\pl T}\cdot A<\intx\sqg R^2>\pr &
\label{state.6}
\end{eqnarray}
We see the entropy is related to the expectation value:\
$A<\intx\sqg R^2>$~ considered in Sec.3, as above.
Using the follwoing results from (\ref{3.18}) ($\xi=1$~is taken),
\begin{eqnarray}
+\mbox{branch solution}\qqq
& \expect=\left\{ \begin{array}{ll}
(8\pi)^2(\frac{2}{w}-\frac{1}{w^2}) & \mbox{for}\ 0<w\leq 1\\
+(8\pi)^2 & \mbox{for}\ 1\leq w
\end{array}
\right. \com & \nn\\
-\mbox{branch solution}\qqq
& \expect=\left\{ \begin{array}{ll}
+(8\pi)^2 & \mbox{for}\ w\leq 1\\
(8\pi)^2(\frac{2}{w}-\frac{1}{w^2}) & \mbox{for}\ 1\leq w
\end{array}
\right. \com & \label{state.7}
\end{eqnarray}
we obtain the expression for the entropy.
\begin{eqnarray}
+\mbox{branch solution}\qqq
& S_{ent}=\left\{ \begin{array}{ll}
-\frac{4\pi}{\ga}(2w-1) & \mbox{for}\ 0<w\leq 1\\
-\frac{4\pi}{\ga} & \mbox{for}\ 1\leq w
\end{array}
\right. \com & \nn\\
-\mbox{branch solution}\qqq
& S_{ent}=\left\{ \begin{array}{ll}
-\frac{4\pi}{\ga} & \mbox{for}\ w\leq 1\\
-\frac{4\pi}{\ga}(2w-1) & \mbox{for}\ 1\leq w
\end{array}
\right. \pr & \label{state.8}
\end{eqnarray}
The graph of $S_{ent}$~ is plotted in Fig.13.
{\vs 5}
\begin{center}
Fig.13\ Entropy per unit mol, $S_{ent}(w)/(4\pi/\ga)$, Pos.Curv.Sol.
\end{center}
The largeness of the absolute value of $S_{ent}$~ in Phase (A) shows
the much amount of freedom of the system, whereas the fixed value
in Phase (B) and (C) shows the possible configurations are
restricted.
$\qq$ From the behaviours of the temperature and the entropy, the cross-over
in the $-$~solution looks to occur
only at one point, $w=1$~, on the w-axis. The corresponding one to
$w_c'$~ in Sec.6 does not appear.
$\qqq$In Fig.14, all phases above are pictorially depicted.
\vspace{10cm}
\begin{center}
Fig.14\ Schematic image of surface in each phase
\end{center}
\section{Discussions and Conclusions}
$\qq$ Among two branches ,the $-$ branch (of the positive curvature) solution
appears in the lattice simulation\cite{ITY1,ITY2}.
It is consistent with the present analysis, where $-$ branch
is energetically prefarable to + branch.
Some features of + branch are the same as those obtained in \cite{KN}
using the conformal field approach\cite{ITY1}. It seems important
to analyse the relation between the present semiclassical approach
and the conformal field approach.
$\qq$ We discuss the meaning and the possible
role of the negative curvature solution.
The existance of the vacua related by the reflection symmetry:\
$R\change -R$\ , is one of stressing points of this paper.
We may say
the appearance of 'dual' solutions reflects the reflection symmetry\ :\
$R\change -R$\ in the 'induced' $R^2$-gravity\
$\Lcal_{ind}=\frac{1}{2\ga}R\frac{1}{\Del}R-\be R^2-\mu$\ .
The symmetry appears manifestly due to the $R^2$-term.
Their topologies, however, are different
:\ the positive curvature solution satisfies
\ $\intx\sqg~R=8\pi$\ , which means the sphere topology, whereas
the negative one satisfies
:\ $\intx\sqg~R=-8\pi\frac{A}{\ep}=-\infty$\ ,\ which means
the toplogy of a sphere with the infinite number of punctures(Fig.9).
We suppose
this reflection symmetry of vacua
is general for manifolds with other topologies.
It means a physical quantity on a manifold can also be calculated on another
manifold with a different topology. It requires further analysis for clarity.
$\qq$ The semiclassical approach can easily provide the physical meaning
such as thermodynamic properties. The pesent system can be regarded as
the closed thermodynamic system where many scalar-matter particles move
in the gravitational potential and whose configuration is thermally
in an equilibrium state. The $R^2$~ coupling, $w$~(or $\be$~), parametrises
the temerature. The phase difference can be thermodynamically
interpreted as the difference of $w$-dependence of the temperature.
$\qq$ The important role of the integration parameter $\la$~ introduced
in (\ref{3.4f}) and of the total derivative term discussed in Sec.5 show
the proper treatment of the area constraint and the topology
constraint is so important to understand the 2d QG. In the treatment,
the infrared regularizations of Fig.2 and of Fig.8 are nicely used.
Evaluation of
the quantum effects to the present classical results is an important
work to be done. It can be taken into account perturbatively as explained
in (\ref{3.4j}). The renormalization has already been analysed in \cite{S1}.
$\qq$ The present approach can be valid for the higher-dimensional quantum
gravity. The 3 dim QG has been recently 'measured' in the Lattice
simulation with a high statistics.
The semiclassical analysis of the data
will soon become an urgent work to be done.
The success of the perturbative 2d QG
using the semiclassical
method is strongly encouraging for the further progress of the perturbative
quantum gravity in the realistic dimensions.
\begin{flushleft}
{\bf Acknowledgement}
\end{flushleft}
The author thanks
N. Ishibashi, H. Kawai, N. Tsuda and T. Yukawa for discussions
and comments about the present work.
{\vs 2}
|
\section{Introduction}
In a series of recent papers \cite{1} Savvidy, Wegner and co-workers
suggested a Gonihedric random surface
action which could be written as
\begin{equation}
S = {1 \over 2} \sum_{<ij>} | \vec X_i - \vec X_j | \theta (\alpha_{ij}),
\label{e4a}
\end{equation}
on triangulated surfaces, where
$\theta(\alpha_{ij}) = | \pi - \alpha_{ij} |^{\zeta}$
with $\zeta$ some exponent
and $\alpha_{ij}$ is the angle between the
embedded neighbouring triangles with common link $<ij>$.
This action is a robust
discretization of the linear size
of a surface, which is a well-defined geometrical
notion that may be constructed in various equivalent
ways. It was intended as an alternative to
gaussian plus extrinsic curvature lattice actions
of the form
\begin{equation}
S = \sum_{<ij>} ( \vec X_i - \vec X_j )^2 +
\lambda \sum_{\Delta_i, \Delta_j} ( 1 - \vec n_i \cdot \vec n_j )
\label{e01}
\end{equation}
which have been much explored \cite{2} as discretizations
of rigid membranes and strings \cite{3}.
Although a simulation showed that the action of equ.(\ref{e4a})
produced flat surfaces \cite{4}, potential problems
arising from the failure to suppress the wanderings of vertices
in the plane were pointed out in \cite{5}
for the action with $\zeta=1$. Possible ways to cure
this are to add additional Gaussian or linear terms to the action
\cite{5a}
or, more satisfactorily, to simply choose $\zeta<1$.
A study of the scaling of the string tension and mass gap
in a dynamically triangulated model with
an additional linear term produced inconclusive results \cite{6},
as have simulations of the gaussian plus extrinsic curvature
action, because of the difficulties of simulating a
complicated action on a dynamical surface.
There is thus some incentive to investigate
alternative approaches to such surface models where
the computational costs are less onerous.
One such approach
for regularizing the Gonihedric action is to
restrict the allowed surfaces to a (hyper)cubic lattice. This
has been pursued in some detail analytically in \cite{7,8,8a}
and one numerical simulation carried out in three dimensions
\cite{8}.
Recently Pietig and Wegner \cite{8b} have demonstrated
with a Peierls contour argument that a transition
{\it does} exist in the three-dimensional case and
obtained a bound on the critical temperature
that is not contradicted by the simulations.
The crucial observation
for this work is that the surface theory
on a cubic or hypercubic lattice can be written equivalently as a
generalized Ising action, where the boundaries
between the spin clusters are the surfaces of the original model.
The latticized Gonihedric model assigns a non-zero action
only to right angled bends in the surface and self-intersections.
Normalizing the weight for a right-angled bend on a link
appropriately leaves
a free parameter $\kappa$ that gives
the relative weight
of a self-intersection of the surface on a link.
The energy of a surface on a cubic
lattice is thus given explicitly by
$E=n_2 + 4 \kappa n_4$, where $n_2$ is the number
of links where two plaquettes meet at a right angle
and $n_4$ is the number of links where four plaquettes
meet at right angles.
A Hamiltonian which reproduces
this energy on a cubic lattice has the form
\begin{equation}
H= 2 \kappa \sum_{<ij>}^{ }\sigma_{i} \sigma_{j} -
\frac{\kappa}{2}\sum_{<<i,j>>}^{ }\sigma_{i} \sigma_{j}+
\frac{1-\kappa}{2}\sum_{[i,j,k,l]}^{ }\sigma_{i} \sigma_{j}\sigma_{k} \sigma_{l}
\label{e1}
\end{equation}
which is of generalized Ising form and contains nearest neighbour ($<i,j>$),
next to nearest neighbour ($<<i,j>>$) and round a plaquette ($[i,j,k,l]$)
terms.
Such actions are not new, having been investigated in some detail using
both mean field methods and simulations in \cite{9}.
However,
the particular
combination of coefficients arising in equ.(\ref{e1}) was not considered
explicitly in this work because it corresponds
to a particularly degenerate set of couplings.
This degeneracy manifests itself in an extended symmetry
in the model:
for any value of $\kappa$
it is possible to flip a plane of spins with no energy penalty,
providing the flipped plane does not intersect any existing
portion of surface.
This gives a vacuum degeneracy reminiscent of a gauge theory,
the difference with a true gauge theory
being that here the symmetries are quasi-global, giving only
the freedom to flip entire planes rather than local spins.
Related surface models have also been
simulated directly in \cite{9a},
but again the set of coefficients
appearing in equ.(3)
was not explored in this work.
A very rich phase structure was observed in \cite{9}, in common with
other Ising models with extended interactions \cite{10}
of various sorts which display first and second order phase boundaries
as well as incommensurate phases.
Given the
generic richness of the phase diagrams
for such generalized Ising models
and the additional symmetries
present in the Gonihedric model, the action of equ.(\ref{e1}) merits investigation
from purely statistical mechanical considerations as well
as from the point of view of finding potential continuum string theories.
In the context of string theory one is looking for
a continuous transition (or transitions) at which
a sensible continuum surface theory may be defined.
It is perhaps worth recalling that
even this does not guarantee
a good continuum surface theory. The interfaces
in the standard nearest neighbour Ising model in
three dimensions,
which has a continuous phase transition,
have been investigated in some detail
recently and found to be very porous objects,
decorated with lots of handles at the scale of the lattice
cutoff \cite{11}. Ideally one might hope that
the surfaces generated by the Gonihedric action were
smoother, given that it is derived from a
sort of stiffness term.
Our motivation in this paper is to investigate
the action of equ.(\ref{e1}) in order to sketch out the
gross features
of its phase structure.
We concentrate on $\kappa=1$ as in \cite{8}, but also discuss other values.
A few
cautionary words are in order before we go on to discuss
the mean field approach and simulations.
As we have noted
the ground state of the action in
equ.(\ref{e1}) is very degenerate as
planes of spins parallel
to any of the cube axes can be flipped at no energy cost.
In the case $\kappa=0$ the degeneracy is even larger
as diagonal planes may be flipped now also.
The ability to flip arbitrary spin planes makes defining a magnetic
order parameter rather
problematic. Even the staggered local order parameters defined
in \cite{9} would miss the lamellar phases with arbitrary
intersheet spacings that could be generated
at no cost by flips of spin planes.
The simulations for $\kappa=1$ in \cite{8} were restricted
to measuring the energy and specific heat as
the exhaustive global order parameters suggested there
\begin{equation}
M^{\mu} = \left< {1 \over L^3} \sum_i \sigma_i^{\mu} (vac) \; \sigma_i \right>
\end{equation}
(where $\sigma_i^{\mu} (vac)$
is one of the possible vacuum
spin configurations with $\mu = 1,2 \ldots 2^{3L}$ for $\kappa=0$
or $\mu = 1,2 \ldots 3 \times 2^L$ for $\kappa \ne 0$ on a lattice of size $L$)
would have been prohibitively slow to measure on even moderately
sized lattices. It is possible to do rather better, however, by making
use of the freedom in choosing boundary conditions on a finite lattice.
It is customary to employ periodic boundary conditions in attempting
to extract critical exponents from a simulation as these tend to
minimize the finite size effects. It is clear that {\it fixed}
boundary conditions in the Gonihedric model would penalize
flipped spin planes by at least a perimeter energy, at the
possible expense of greater finite size effects. A quick test
simulation shows that such boundary conditions do
pick out the purely ferromagnetic ground state from the
many equivalent possibilities and allow the measurement
of the simple ferromagnetic order parameter
\begin{equation}
M = \left< {1 \over L^3} \sum_i \sigma_i \right>.
\end{equation}
One can, in fact, have the best of both worlds by continuing
to employ periodic boundary conditions to reduce the finite size
effects whilst fixing any two perpendicular planes
of spins to pick out the ferromagnetic ground state.
The fixed planes of spin are, in effect, more akin
to a gauge fixing of the spin flip symmetry
than boundary conditions {\it per se}.
In the simulations reported later in this paper
we employed three fixed perpendicular planes
of spins as a safety measure, with essentially
identical results.
\section{Zero Temperature and Mean Field}
As the Gonihedric model is
a special case of the general action considered in \cite{9}
we can apply the methods used there
for both the zero temperature phase diagram and mean field
theory.
For the zero temperature
case this involves writing the full lattice
Hamiltonian as a sum over individual cube Hamiltonians
\begin{equation}
h_c = \frac{\kappa}{2}\sum_{<i,j>} \sigma_{i} \sigma_{j} - \frac{\kappa}{4} \sum_{<<i,j>> }\sigma_{i} \sigma_{j}
+ \frac{1-\kappa}{4} \sum_{[i,j,k,l]}\sigma_{i} \sigma_{j} \sigma_{k} \sigma_{l}
\end{equation}
and observing that if the lattice can be tiled by
a cube configuration minimizing the individual $h_c$
then the ground state energy density is
$\epsilon_0 = min\; h_c$.
We list the inequivalent spin
configurations on a single cube and their
multiplicities in Table.1 using the same notation
as \cite{9} but with our choice of couplings
to highlight the degeneracies that appear with the Gonihedric
action.
In the list of spins the first column represents one face of the cube
and the second the other.
In the table two configurations are considered equivalent if one can be transformed
into the other by reflections and rotations
or if they are related by a global spin flip. The antiferromagnetic image
of a configuration
is obtained by flipping the three nearest neighbours and the spin
at the other end of the cube diagonal from a given spin and is denoted by
an overbar.
With the Gonihedric values of the couplings the freedom to flip spin planes is clear
even at this level as $\psi_0$, which would represent a ferromagnetic state
when used to tile the lattice, and $\psi_6$ which would represent flipped
spin layers,
have the same energy for any value of $\kappa$. The higher energy configurations
$\psi_4$ and $\psi_{\bar 4}$ are also identical.
The degeneracies increase when
$\kappa=0$, the club
of
states of energy $-3/2$ is now composed of $\psi_0, \psi_{\bar 0}, \psi_6, \psi_{\bar 6}$
and various extra degeneracies appear for higher
energy states. The new ground states $\psi_{\bar 0}, \psi_{\bar 6}$
reflect the additional freedom to flip diagonal planes of spins that is present at
$\kappa=0$.
In the mean field approximation the spins
are in effect replaced by the average site magnetizations.
The calculation
of the mean field free energy is an elaboration of the
method used
above to investigate the ground states
in which the energy is decomposed into a sum of individual cube terms.
The next to nearest neighbour and plaquette interactions
in the Gonihedric model give
the total mean field
free energy as the sum of elementary cube free energies $\phi(m_{c})$, given by
\begin{equation}
\phi{(m_{C})}=- \frac{\kappa}{2}\sum_{<i,j>\subset C} m_{i} m_{j} + \frac{\kappa}{4} \sum_{<<i,j>>\subset C }m_{i} m_{j} \]
\[ - \frac{1-\kappa}{4} \sum_{[i,j,k,l] \subset C}m_{i} m_{j} m_{k} m_{l} + \frac{1}{16}
\sum_{i \subset C}[(1+m_{i})ln(1+m_{i})+ (1-m_{i})ln(1-m_{i})]
\end{equation}
where $m_{C}$ is the set of the eight magnetizations of the elementary cube.
This gives a set of eight mean-field equations
\begin{equation}
\frac{\partial\phi(m_{C})}{ \partial m_{i}}_{(i=1 {\ldots} 8)} =0
\end{equation}
(one for each corner of the cube)
rather than the familiar single equation for the standard nearest neighbour Ising
action.
More explicitly, we have
\begin{eqnarray}
m_{1}&=& tanh[4\beta \kappa(m_{2}+m_{4}+m_{5})-2\beta \kappa(m_{3}+m_{6}+m_{8}) \nonumber \\
& & + 2\beta(1- \kappa)(m_{2}m_{3}m_{4}+ m_{2}m_{5}m_{6}+m_{4}m_{5}m_{8}) ] \nonumber \\
& \vdots& \nonumber \\
m_{8}&=& tanh[4\beta \kappa(m_{2}+m_{5}+m_{7})-2\beta \kappa(m_{1}+m_{3}+m_{6}) \nonumber \\
& & + 2\beta(1-\kappa)(m_{3}m_{4}m_{7}+m_{1}m_{4}m_{5}+m_{5}m_{6}m_{7})]
\label{e2a}
\end{eqnarray}
where we have labelled the magnetizations on a face of the cube counterclockwise $1 \ldots 4$
and similarly for the opposing face $5 \ldots 8$
as shown in Figure.1.
If we solve these equations iteratively we arrive at
zeroes for a paramagnetic phase or various combinations
of $\pm 1$ for the magnetized phases on the
eight cube vertices, and the mean field ground state
is then give by gluing together the elementary cubes consistently
to tile the complete lattice, in the manner
of the ground state discussion.
Turning loose a numerical solver on the mean field equs.(\ref{e2a}) gives generically a
single transition to one of the phases
listed in Table.1 from the high temperature
paramagnetic phase. The transition temperatures and the resulting low temperature
phase are listed in Table.2.
We have taken the liberty of carrying out global
flips where necessary to tidy up the table.
Rather remarkably, we see that apart from $\kappa=0$
the transition appears to be to the simple ferromagnetic phase, $\psi_0$. However, remembering
that $\psi_0$ and $\psi_6$ have the same energy the best we can say is that
we end up in a layered phase with arbitrary interlayer spacing in all directions.
Although the $\kappa=0$ case appears to be superficially different,
the $\psi_{\bar 0}$ phase that is found at low temperature here
is one of the phases that is degenerate with $\psi_0$ and $\psi_6$
when $\kappa=0$.
Although $\kappa=1$ fits the pattern as far as
a transition to $\psi_{0,6}$ at decreasing $\beta$
is concerned it appears to be rather atypical in that
further transitions are observed at larger $\beta$.
However, this is a numerical instability
that is peculiar to this particular value of $\kappa$.
It was observed in \cite{9} that an iterative solution
of the mean field equations
written in the form
\begin{equation}
m_i^{(n+1)} = f[E_{,i} (m^n)]
\end{equation}
where $E$ is the individual cube Hamiltonian
could fail to converge if an eigenvalue of
$ \partial m^{(n+1)}_i / \partial m^n_j$
was less than $-1$. Modifying the equations
to
\begin{equation}
m_i^{(n+1)} = { \left( f[E_{,i} (m^n)] + \alpha m^n_i \right) \over 1 + \alpha}
\end{equation}
for suitable $\alpha$ cures this. This is precisely what happens
for $\kappa=1$, where introducing a non-zero $\alpha$ suppresses the
extra ``transitions''.
In summary, the mean field theory suggests a rather simple
phase diagram for the Gonihedric model
with action equ.(\ref{e1}), with a single transition
from a paramagnetic phase
to a degenerate ``layered'' phase
that is pushed down to $\beta=0$ at large $\kappa$.
The low temperature phase is generically
of the $\psi_{0,6}$ type, apart from $\kappa=0$ where
we see a $\psi_{\bar 0, \bar 6}$ phase that is degenerate
with these. The degeneracy of the
ground states that are indicated by these results
are, as they should be,
consistent with the symmetries of the original
full action.
We now go on
to see how the zero-temperature and mean field results
tally with a direct Monte-Carlo
simulation.
\section{Simulations}
We carried out simulations
with $\kappa=1$ on lattices
of size $10^3,12^3,15^3,18^3,20^3$ and $25^3$
and for $\kappa=0,2,5,10$
on lattices of size $10^3,15^3,20^3$ and $25^3$.
Unless stated otherwise periodic boundary conditions were imposed
in the three directions and three internal perpendicular planes
of spins fixed to be $+ 1$.
We carried out 50K
sweeps for most $\beta$ values,
increasing to 500K sweeps near the observed
phase transition point. Measurements were carried out
every sweep after allowing sufficient time
for thermalization.
A simple Metropolis update was used because of the difficulty
in concocting a cluster algorithm for a Hamiltonian with
such complicated interaction terms. The program was tested
on the standard nearest neighbour Ising model and
the some of the parameters used in
the generalized Ising models of \cite{9} to ensure it was working.
We measured the usual thermodynamic quantities for the model:
the energy $E$, specific heat $C$,
(standard) magnetization $M$, susceptibility $\chi$
and various cumulants. As the large $\beta$ limit of the
energy should be determined by the zero-temperature
analysis of the preceding section, we consider
the energy first.
The absolute value of the energy for various $\kappa$
on lattices of size $L=20$ is plotted against $\beta$
in Figure.2, where we can see that the zero temperature
prediction of $3(1+ \kappa) / 2$ is satisfied
with good accuracy for sufficiently large $\beta$.
We can therefore observe that the zero-temperature/mean-field
analysis has correctly identified the ground state(s) of the theory:
$\psi_{0,6}$ for $\kappa \ne 0$; or $\psi_{0,\bar 0,6,\bar 6}$ for
$\kappa=0$ as these are the only states with the observed
energies. Having satisfied ourselves that the
simulation is finding the correct ground state energy, we
can go on to consider extracting some of the critical exponents for the transition.
In what follows we will, as advertized, discuss in some detail the case $\kappa=1$
before commenting more briefly on the other
values of $\kappa$.
With only periodic
boundary conditions the possibility of the simple ferromagnetic
ordered state $\psi_0$ can be excluded by looking at
the magnetization $M$, which for all
$\kappa$ is either zero or fluctuates wildly as $\beta$
is changed, reflecting the freedom to flip spin planes.
Curtailing this freedom by fixing the three perpendicular
spin planes picks out the transition
to a simple ferromagnetic ground state
and the low temperature limit
of the magnetization becomes
one for all $\kappa$.
The magnetization cumulant
\begin{equation}
U_M = { <M^4> \over <M^2>^2 }
\end{equation}
is well defined once the three spin planes are fixed,
but it does not show the standard behaviour
of a low $\beta$ limit of three and a high
$\beta$ limit of one, asymptoting to a value slightly
larger than one at low $\beta$, as can be seen in Figure.3.
This is because the fixed planes still leave a residual
magnetization at low $\beta$, which is sufficient
to make this limit look magnetized for the sizes
of lattice we simulate. Nonetheless, it is still
possible to apply the usual scaling analysis in the critical region,
and the crossing of the cumulant plots for different latttice
sizes gives an estimate of $\beta_c = 0.44(1)$ for the critical
temperature, which is in good agreement with the value
reported in \cite{8} that was extracted by looking
at the change in behaviour of the spin/spin correlation
as $\beta_c$ was approached. As noted in \cite{8}
this $\beta_c$ is very close (in our case
within the error bars) to that of the standard
two-dimensional Ising model on a square lattice.
The rather sharp nature of the crossing, or more accurately
collapse down to a single line, makes it difficult to extract
a value for $\nu$ from the ratio of the slopes
of the Binder's cumulant curves at the
critical point, so we take a different tack and
consider the scaling of the maximum slope,
which we would also expect to scale as
$L^{1 / \nu}$ \footnote{This tactic
works well in, for instance, simulations of Ising
models coupled to two-dimensional quantum gravity.}.
This gives an estimate of $\nu = 1.2(1)$.
We can also look at both
the finite size scaling and direct fits to
the susceptibility $\chi$, namely
$\chi = A L^{\gamma \over \nu}$ and $\chi = \tilde A | \beta - \beta_c | ^{\gamma}$,
to attempt to extract $\nu$. We choose this in preference to the
specific heat fits $C = B + D L^{\alpha \over \nu}$ and $C = \tilde B + \tilde D | \beta - \beta_c |^{\alpha}$
because of the absence of an adjustable constant.
The susceptibility data is plotted in Figure.4, and shows a clear peak. This should be contrasted
with the case of no fixed planes where the $\beta>\beta_c$ region is rendered meaningless noise by
the lack of a well-defined magnetization for the myriad of ground states.
We find the finite size scaling fit gives $\gamma / \nu = 1.79(4)$ with
a high quality, and feeding the critical value of $\beta_c=0.44$ into the direct fit
on the $L=25$ lattice gives $\gamma=1.60(2)$ with rather lower quality. The deduced
value for $\nu$ is thus $1.10(5)$. These fits give values
close to those for the standard two-dimensional Ising model
with only nearest neighbour interactions, where we have $\gamma=1.75, \nu =1$.
As a consistency check on our values of $\beta_c$ and $\nu$, we plot
the $\beta$ values where the specific heat peaks and the $\beta$ values
where the maximum slope of the Binder's cumulant curve occurs,
both of which can serve as estimates of the pseudocritical
temperature on finite size lattices, against
$L^{-{1 \over \nu}}$. We would expect this to be a straight line with
intercept $\beta_c$. The plot is shown in Figure.5 for the
choice $\nu=1$ with other
values in this neighbourhood giving essentially identical results.
We can see that the estimate of $\beta_c=0.437(7)$
coming from the two possibilities is both self-consistent and in agreement
with the value obtained from cumulant crossing.
The above, apparently consistent, set of results presents us with something of a
dilemma when it comes to the analysis of the specific heat peak, which is shown in
Figure.6. The hyperscaling relation $\alpha = 2 - \nu d$ indicates that, if a
value of $\nu \simeq 1$ is to be believed, the specific heat should display a
cusp ($\alpha = 2 - \nu d \simeq -1$) rather than a divergence. This does not appear
to be the case for the data in the figure.
Setting aside the hyperscaling result for the moment
and performing direct power law fits to
$C = B + D L^{\alpha \over \nu}$ and
$C = \tilde B + \tilde D | \beta - \beta_c |^{\alpha}$ gives poor fits.
A logarithmic fit of the form
$C = B + D \log ( L )$ or $C = \tilde B + \tilde D \log (\beta - \beta_c)$
gives much better, but still not particularly good, results
so the evidence is inconclusive.
Another line of attack
for obtaining an estimate of $\alpha$ is to use the finite size
scaling of the energy itself
$E \simeq E_0 + E_1 L^{\alpha -1 \over \nu}$. With
direct measurements in this form one gains nothing
in general over the specific heat
fits as there is still a regular term $E_0$
to be dealt with, although for models with higher than
second order transitions this approach may be preferable \cite{wh}.
However, if one has measurements for two different sets
of boundary conditions available the regular term would be
expected to be the same for both and the energy
difference could be used for a simple power law fit
to extract $( \alpha -1 ) / \nu$. We are currently
measuring the string tension in the Savvidy model using
two sets of fixed boundary conditions \cite{esp}, one with
all positive spins and one with half positive and half negative
spins \footnote{It is not possible to use antiperiodic
boundary conditions in the Savvidy model to enforce
an interface because of the plane flip symmetry. Something
more coercive, in the form of these fixed boundary conditions is
required.}. For these measurements we would expect
\begin{equation}
\Delta E = E_{++} - E_{+-} = A L^{\alpha - 1 \over \nu}
\end{equation}
where $E_{++}$ is the energy for all positive spins and
$E_{+-}$ is the energy for half positive and half negative
spins.
A fit gives $( \alpha - 1 ) / \nu = -1.3(2)$, which is still
marginally consistent with $\alpha=0$.
There are two possible interpretations of the results. The first is that
the value of $\nu$ measured is simply wrong, not inconceivable as it
appears either as a derived quantity from
the slope of the cumulant or from the two fits
to the susceptibility rather than being
measured directly.
However, a second possible interpretation
is to accept the fits to $\nu$ at face value and posit
that the model sees an effective dimension of $d=2$
in order to avoid violating the hyperscaling relation.
In this case we could recover the full set of
two-dimensional Ising model exponents.
Although this would be highly unusual, it should be remembered
that the energy in the model is essentially linear in form
rather than being an area, so there is some resemblance to
a two-dimensional model.
A direct fit to the magnetization exponent on the largest
lattice size $M \simeq |\beta - \beta_c|^{\beta}$
(with apologies for the profusion of betas!) with $\beta_c$
fixed to be $0.44$ gives $\beta = 0.12(1)$, but the quality
is low, whereas a finite size scaling fit gives a
much lower value of $\beta / \nu = 0.04(1)$. It is
possible that the fixed spin planes, whose residual
magnetization we have not allowed for in the fits,
are biasing the finite size scaling fit, but intuitively
one would expect their
effects (of order $3 / L$) to increase
rather than decrease the estimated exponent
by pushing up the measured magnetization on the smaller lattices.
We now discuss more briefly the other $\kappa$ values
that were simulated, namely $\kappa=0,2,5,10$.
Firstly we can note that in qualitative
terms the transition appears similar to the $\kappa=1$ case,
a not entirely trivial result as $\kappa \neq 1$ introduces
round a plaquette interactions that are missing for $\kappa=1$.
With the periodic plus fixed plane boundary conditions
we still have peaks in the susceptibility and specific heat
and a ferromagnetically ordered phase appearing
at low temperature.
The mean field
analysis suggests that as $\kappa$ is increased
$\beta_c$ should drop sharply. This is {\it not} observed
in the simulations, the crossing of the Binder's
cumulants indicating no change within the error
bars for the estimates of $\beta_c$,
giving $\beta_c = 0.44(1)$
from $\kappa=1$ to $\kappa=10$.
There is a sharper difference with the $\kappa=0$ results
which show a crossing at $\beta_c = 0.50(1)$.
In general
mean field theory will underestimate $\beta_c$, so the
measured results are in agreement with this
and do not contradict the bound obtained in
\cite{8b}.
The similarity of the transitions
for different $\kappa$ extends beyond $\beta_c$. The
measurements of $\gamma / \nu$ listed in Table.3 below for
all the non-zero $\kappa$
suggest that the critical behaviour is unchanged
by varying $\kappa$.
\begin{center}
\begin{tabular}{|c|c|c|c|c|} \hline
$\kappa$ & $1$ & $2$ & $5$ & $10$ \\[.05in]
\hline
$\gamma / \nu$ & $1.79(4)$ & $1.6(1)$ & $1.9(1)$ & $1.75(6)$ \\[.05in]
\hline
\end{tabular}
\end{center}
\vspace{.1in}
\centerline{Table 3: Fits to $\gamma / \nu$ for the non-zero $\kappa$ values.}
\noindent
The specific heat curves and magnetization present a similar story,
with all looking similar to the $\kappa=1$ case.
From this evidence it would seem that varying $\kappa$,
at least for $\kappa \ge 1$,
gives little if any change in the exponents and transition
temperature.
The story is slightly different for $\kappa=0$.
As we have already noted $\beta_c=0.50(1)$,
and other differences are apparent.
Without the fixed spin planes (ie with
only periodic boundary conditions) the
susceptibility when $\beta<\beta_c$ for
non-zero $\kappa$ values is similar to the fixed plane
case described above and becomes meaningless
when $\beta>\beta_c$ where the magnetization is ill-defined.
The $\kappa=0$ model presents qualitatively different behaviour
in that the susceptibility is one for
$\beta<\beta_c (\simeq 0.5)$ and zero for $\beta>\beta_c$.
However, this step
function behaviour disappears when the fixed
spin planes are introduced and we recover a divergent peak
as for the other $\kappa$ values. The phase structure for $\kappa=0$
with the fixed spin planes
also appears to be broadly similar to other $\kappa$, giving
a single transition to a low temperature magnetized phase.
It would be interesting to examine in detail values of
$\kappa$ between zero and one to see if there was a
smooth change in, for example, $\beta_c$
as $\kappa \rightarrow 0$. This would give some indication
of whether $\kappa=0$ really was a special point,
or joined on smoothly to the continuum of non-zero values.
A test simulation at $\kappa=0.5$ still gives
very similar results to $\kappa=1$, for example.
\section{Conclusions}
We have conducted zero-temperature, mean-field and Monte-Carlo
investigations of the generalized Ising model action suggested
in \cite{7,8,8a} as a cubic lattice discretization
of the Gonihedric string action \cite{1}
using essentially the methods of \cite{9}. Although the
phase structure of such generalized Ising models is
generically very rich \cite{10}, the one parameter family
of models examined here seems to be a fairly
simple ``slice'' of the phase diagram, with one transition
to a layered ground state
when periodic boundary conditions are imposed.
This degenerate layered state is a consequence of
the plane spin flip symmetry that is present in the model
for all $\kappa$, but a judicious choice of boundary conditions
- fixing enough perpendicular spin planes - allowed us
to pick out an equivalent ferromagnetic ground state
for the purposes of simulations.
The zero-temperature/mean-field analyses are in agreement
with the Monte-Carlo simulations on the nature of the ground
state and its energy, but the simulations
indicate a transition temperature that changes little,
if at all, from its value at $\kappa=1$
($\beta_c \simeq 0.44$) for other non-zero $\kappa$ values.
The mean field theory on the other hand gives
a fairly sharp decline in $\beta_c$ as $\kappa$ is increased.
The simulations at $\kappa=1$
indicate that the fitted exponents,
with the exception of the finite size scaling fit to
$\beta / \nu$,
and even the critical temperature are all in
the vicinity of those for the two-dimensional Ising model,
though given our modest statistics
it would be foolhardy to claim they were identical on
the basis of the current fits.
Comparison with the other non-zero $\kappa$ values
also gives similar exponents and critical temperatures.
There is some evidence
that the $\kappa=0$ model is a special case: -
in the zero temperature and mean field analyses
more ground states are allowed and in the simulations
a different critical temperature is observed
and the behaviour of the susceptibility is radically different
when no spin planes are fixed.
An immediate extension of the current work, given the closeness
of the fitted exponents to the two-dimensional Ising model,
is to carry out a higher statistics simulation
near the transition point in order to pin down
the various exponents and $\beta_c$ more accurately.
A further test of the critical behaviour of the model
would be to investigate the scaling of the string tension
as one approached the critical point, along the lines of \cite{esp}.
The various higher dimensional
generalizations that were formulated in \cite{7,8,8a}
also merit investigation.
If we return to our original stringy motivation
it would be useful to characterize the surfaces generated
by the Gonihedric action in the style of \cite{11}
to see whether they were any less ``spongy'' than those
in the standard 3D Ising model.
As a playground for exploring plaquette discretizations
of string and gravity inspired models the generalized
Ising models clearly have some interesting quirks
that are worthy of further exploration. It would certainly
be amusing to show that a candidate discretized string model
was a close relation of the {\it two}-dimensional
Ising model.
\section{Acknowledgements}
R.P.K.C. Malmini was supported by Commonwealth Scholarship
SR0014.
\vfill
\eject
|
\section{Introduction}
The cooling history of a newly born neutron star in the center
of a supernova (SN) is mainly determined by neutrino diffusion.
Numerical simulations employing the lowest order neutrino
interaction rates calculated within the Glashow-Salam-Weinberg
theory predict a cooling time scale which agrees remarkably well
with the neutrino signal observed from SN 1987 A~\cite{Schramm}.
The emission of novel weakly interacting particles like
axions~\cite{R3}
could change the cooling time scale substantially which in turn
allows to derive constraints on the properties of such
particles~\cite{Burrows}.
Within linear response theory weak interaction rates
with a medium of nonrelativistic nucleons are determined, apart
from the weak phase space, by only two dynamical structure
functions, one for the density and one for the nucleon
spin-density~\cite{Iwamoto1,R2}. Some work has
been devoted to their calculation but either the Landau theory
of quasiparticles was applied assuming
a ``cold'' nuclear medium~\cite{Iwamoto1,cold} or the authors
focused on quasielastic scattering studying static structure
functions~\cite{Iwamoto1,Sawyer,hot}.
Interactions of neutrinos and axions with a nonrelativistic
nuclear medium are mainly governed by the local nucleon spin-density and its
fluctuations. To lowest nontrivial order in the spin dependent
nucleon-nucleon interactions causing these fluctuations, the
relevant weak processes are of the nucleon bremsstrahlung type.
Due to the Landau
Pomeranchuk Migdal (LPM) effect~\cite{LPM} which accounts for multiple
nucleon scattering the inelasticity of these processes depends
on the nucleon spin flip rate. In addition, once this rate
becomes
considerably larger than the medium temperature $T$, the total
weak interaction rates tend to be suppressed~\cite{R1}.
Since perturbative estimates for the nucleon spin flip rate
can be as high as $\simeq50T$ around nuclear densities, this
could have profound implications for SN core
physics~\cite{R2,R1,Keil,Janka}.
By dramatically reducing the predicted SN
cooling time scale it would spoil the agreement between theory
and the observed neutrino pulse from SN 1987 A~\cite{Keil}. On
these phenomenological grounds it has been suggested that
axial-vector neutrino scattering cross sections might be roughly density
independent~\cite{Janka} instead of being suppressed at high
densities by the LPM effect.
In this letter we derive a new sum rule for the dynamical spin-density
structure function (SSF) which provides an independent
theoretical argument supporting this conjecture. It also predicts
that emissivities for weakly interacting particles should
increase somewhat slower than the lowest order rates at high
densities.
\section{The Spin-Density Structure Function}
In terms of the nucleon field operator in the nonrelativistic
limit, $\psi(x)$, the spin-density operator is given by
${\hbox{\boldmath $\sigma$}}(x)={1\over2}\psi^\dagger(x)
{\hbox{\boldmath $\tau$}}\psi(x)$ where
${\hbox{\boldmath $\tau$}}$ are the
Pauli matrices. In the following we denote the momentum,
coordinate, and spin operators for a single nucleon by ${\bf p}_i$,
${\bf r}_i$, and ${\hbox{\boldmath $\sigma$}}_i$, respectively, where
$i=1,\cdots,N_b$ runs
over $N_b$ nucleons. Then, for a normalization volume $V$, we
can define the Fourier transform
\begin{equation}
{\hbox{\boldmath $\sigma$}}(t,{\bf k})={1\over V}\int d^3{\bf r}
e^{-i{\bf k}\cdot{\bf r}}{\hbox{\boldmath $\sigma$}}(t,{\bf r})=
{1\over V}\sum_{i=1}^{N_b}e^{-i{\bf k}\cdot{\bf r}_i}
{\hbox{\boldmath $\sigma$}}_i\,.\label{sk}
\end{equation}
In terms of these operators and the baryon density $n_b$
the SSF is defined
as~\cite{R2,Janka}
\begin{equation}
S_\sigma(\omega,{\bf k})={4\over3n_b}\int_{-\infty}^{+\infty}
dte^{i\omega t}\left\langle{\hbox{\boldmath $\sigma$}}(t,{\bf k})
\cdot{\hbox{\boldmath $\sigma$}}(0,-{\bf k})
\right\rangle\,,\label{sdef}
\end{equation}
where $(\omega,{\bf k})$ is the four-momentum transfer to the
medium. The expectation value $\langle\cdots\rangle$ in
Eq.~(\ref{sdef}) is taken over a thermal ensemble.
The contribution of $S_\sigma$ to the neutrino scattering rate (per
final state density) from four momentum $(\omega_1,{\bf k}_1)$ to
$(\omega_2,{\bf k}_2)$ can be written as ${1\over4}G_{\rm
F}^2C_A^2n_b(3-\cos\,\theta)S_\sigma(\omega_1-\omega_2,{\bf k}_1-{\bf k}_2)$
with $G_{\rm F}$ the Fermi constant, $C_A$ the relevant
axial-vector charge, and $\theta$ the angle between ${\bf k}_1$
and ${\bf k}_2$. Similarly, the rate for pair production
would read
${1\over4}G_{\rm F}^2C_A^2n_b(3+\cos\,\theta)S_\sigma(-\omega_1-\omega_2,
-{\bf k}_1-{\bf k}_2)$~\cite{R2}. The axion emission rate per volume,
$Q_a$, is governed by the same structure function [in an
isotropic medium $S_\sigma(\omega,{\bf k})=S_\sigma(\omega,k)$ only
depends on $k=\vert{\bf k}\vert$]:
\begin{equation}
Q_a={C_N^2n_b\over(4\pi)^2f_a^2}\int_0^\infty d\omega\,\omega^4
S_\sigma(-\omega,\omega)\,.\label{Qa}
\end{equation}
Here, $f_a$ is the Peccei-Quinn scale and the numerical factor
$C_N$ depends on the specific axion model~\cite{R3}. Neutrino
opacities and axion emissivities are therfore mainly determined
by the SSF at thermal energies $\omega\simeq k\mathrel{\mathpalette\fun <} T$.
Eq.~(\ref{sdef}) implies
\begin{equation}
\int_{-\infty}^{+\infty}{d\omega\over2\pi}\,\omega S_\sigma(\omega,{\bf k})
=-{4\over3n_b}\left\langle[H,{\hbox{\boldmath
$\sigma$}}(0,{\bf k})]\cdot{\hbox{\boldmath $\sigma$}}(0,-{\bf k})
\right\rangle\,,\label{sum1}
\end{equation}
where $H$ is the Hamiltonian of the system of interacting
nucleons for which we assume the following form:
\begin{equation}
H=H_0+H_{\rm int}=\sum_{i=1}^{N_b}{{\bf p}_i^2\over2M}+{1\over2}
\sum_{i\neq j}^{N_b}V({\bf r}_{ij},{\hbox{\boldmath $\sigma$}}_i,
{\hbox{\boldmath $\sigma$}}_j)\,.\label{H}
\end{equation}
Here, ${\bf r}_{ij}={\bf r}_i-{\bf r}_j$, $M$ is the
free nucleon mass, and $V({\bf r}_{ij},{\hbox{\boldmath
$\sigma$}}_i,{\hbox{\boldmath $\sigma$}}_j)$ is the spin
dependent two nucleon interaction potential. For notational
simplicity we restrict ourselves to only one nucleon species for
the moment; the general case will be discussed further below.
For free nucleons one gets
$\int_{-\infty}^{+\infty}(d\omega/2\pi)\,\omega
S_\sigma(\omega,{\bf k})={\bf k}^2/2M$, in analogy to
the well known f sum rule for the dynamical density structure
function. In the latter case nucleon number conservation ensures
that the f sum rule even holds in the
presence of velocity independent interactions. In contrast, the
f sum for the SSF is modified in the presence of spin dependent
interactions since the nucleon spin is in general not conserved.
For one nucleon species the most general two nucleon interaction
potential is of the form ~\cite{ST}
\begin{equation}
V({\bf r},{\hbox{\boldmath $\sigma$}}_1,{\hbox{\boldmath
$\sigma$}}_2)=U(r)+U_S(r){\hbox{\boldmath $\sigma$}}_1
\cdot{\hbox{\boldmath $\sigma$}}_2
+U_T(r)\left(3{\hbox{\boldmath $\sigma$}}_1\cdot\hat{\bf r}\,
{\hbox{\boldmath $\sigma$}}_2\cdot\hat{\bf r}-{\hbox{\boldmath
$\sigma$}}_1\cdot{\hbox{\boldmath $\sigma$}}_2\right)
\,,\label{Vint}
\end{equation}
where ${\bf r}={\bf r}_{12}$, $r=\vert{\bf r}\vert$, and $\hat{\bf
r}={\bf r}/r$. We denote the spin dependent terms by
$V^{S}_{ij}=U_S(r_{ij}){\hbox{\boldmath $\sigma$}}_i
\cdot{\hbox{\boldmath $\sigma$}}_j$ (the ``scalar force'') and
$V^{T}_{ij}=U_T(r_{ij})\left(3{\hbox{\boldmath
$\sigma$}}_i\cdot\hat{\bf r}_{ij}\,{\hbox{\boldmath $\sigma$}}_j\cdot
\hat{\bf r}_{ij}-{\hbox{\boldmath $\sigma$}}_i
\cdot{\hbox{\boldmath $\sigma$}}_j\right)$ (the ``tensor
force''). In order to
calculate the additional commutator in Eq.~(\ref{sum1})
from Eqs.~(\ref{sk}), (\ref{H}) and (\ref{Vint}) we make use of
the commutation relations
$\left[\sigma_i^a,\sigma_j^b\right]=
i\delta_{ij}\epsilon^{abc}\sigma_i^c$,
where $i,j=1,\cdots,N_b$ and $\epsilon^{abc}$ is the total
antisymmetric tensor in the spatial indices $a,b,c$.
After some algebra and using the symmetry
properties of the Hamiltonian the modified sum rule reads
\begin{equation}
\int_{-\infty}^{+\infty}{d\omega\over2\pi}\,\omega
S_\sigma(\omega,{\bf k})
={{\bf k}^2\over2M}\label{sum1a}
-{4\over3N_b}\sum_{i\neq j}^{N_b}
\left\langle V^{S}_{ij}+V^{T}_{ij}+\cos\,{\bf k}\cdot{\bf r}_{ij}
\left({1\over2}V^{T}_{ij}-V^{S}_{ij}\right)
\right\rangle\,.
\end{equation}
The kinetic nucleon recoil term is in general
negligible compared to the $V$-dependent terms which govern the
inelasticity of axial-vector interactions.
If $V({\bf r},{\hbox{\boldmath $\sigma$}}_1,{\hbox{\boldmath
$\sigma$}}_2)\mathrel{\mathpalette\fun >}-\alpha/r^s$ with $\alpha>0$ and $s<2$
the eigenvalues of $H$ are bounded from below and the r.h.s. of
Eq.~(\ref{sum1a}) is finite as long as $U_S(r)$ and $U_T(r)$
are integrable. This is the case for typical meson exchange
potentials with hard core repulsion~\cite{ST,BD,FM}.
Assuming the three terms in Eq.~(\ref{Vint}) to be of similar
size the r.h.s. of Eq.~(\ref{sum1a}) is roughly proportional to
the average interaction energy per nucleon $W$. At zero temperature
and for SN core densities and compositions,
$W\simeq30\,{\rm MeV}$ corresponding to an average binding energy
of about $10\,{\rm MeV}$ per nucleon. For $T\ga10\,{\rm MeV}$
nucleons are bound more weakly and $W$ should be considerably
smaller. We can therefore write
\begin{equation}
\int_{-\infty}^{+\infty}{d\omega\over2\pi}\,\omega
S_\sigma(\omega,{\bf k})\simeq4W
\la100\,{\rm MeV}\,,\label{Vmax}
\end{equation}
where the inequality is a conservative bound reflecting our poor
knowledge about the equation of state for hot nuclear matter.
Since it involves bound state energies, Eq.~(\ref{Vmax}) is a
nonperturbative result and will play an
important role for the high density behavior of weak interaction
rates below.
The dependence on the momentum transfer ${\bf k}$ in
Eq.~(\ref{sum1a}) is expected to be only modest.
In fact, for $k\mathrel{\mathpalette\fun <} T\la50\,{\rm MeV}$, we
have $\vert{\bf
k}\cdot{\bf r}\vert\ll1$ within the range of the potential
$r_s\simeq1/m_\pi$ which is determined by the pion mass
$m_\pi\simeq140\,{\rm MeV}$. We can thus go to the long
wavelength limit~\cite{Iwamoto1,R2,Sawyer,Keil,Janka},
${\bf k}\rightarrow0$, using
$S_\sigma(\omega)\equiv S_\sigma(\omega,{\bf k}\rightarrow0)$.
Eq.~(\ref{sum1a}) then simplifies to
\begin{equation}
\int_{-\infty}^{+\infty}{d\omega\over2\pi}\,\omega S_\sigma(\omega)
=-{4\over N_b}\langle H_T\rangle
\,,\label{sum1b}
\end{equation}
where $H_T={1\over2}\sum_{i\neq j}^{N_b}V^{T}_{ij}$.
First, note that the scalar force does not contribute to
Eq.~(\ref{sum1b}) because it
conserves the total nucleon spin ${\hbox{\boldmath
$\sigma$}}(0,{\bf k}\rightarrow0)$
[see Eq.~(\ref{sk})]. Below nuclear densities the
nucleon-nucleon ($NN$) interaction is dominated by one-pion exchange
(OPE) leading to a tensor force.
This contribution induces a spin orbit coupling and does therefore not
conserve the total nucleon spin. Thus only the tensor force
contributes to Eq.~(\ref{sum1}) in the long
wavelength limit. This agrees with the lowest order
bremsstrahlung calculation for ${\bf k}\rightarrow0$~\cite{FM}.
Finally, note that the r.h.s. of Eq.~(\ref{sum1b}) is
positive as it should be since the interaction induced
correlations reduce $\langle H_T\rangle$ below the value for
free nucleons, $\langle H_T\rangle=0$.
An additional sum rule~\cite{R2,Keil,Janka} can be
obtained by integrating Eq.~(\ref{sdef}) and using
Eq.~(\ref{sk}) in the long wavelength limit:
\begin{equation}
\int_{-\infty}^{+\infty}{d\omega\over2\pi}\,S_\sigma(\omega)
=1+{4\over3N_b}\left\langle\sum_{i\neq j}^{N_b}{\hbox{\boldmath
$\sigma$}}_i\cdot{\hbox{\boldmath $\sigma$}}_j\right\rangle
\,.\label{sum2}
\end{equation}
Note that for free nucleons Eqs.~(\ref{sum1b}) and (\ref{sum2})
yield $S_\sigma(\omega)=2\pi\delta(\omega)$, whence only elastic
scattering on the medium is possible in the absence of $NN$
interactions.
\section{Dilute Medium Limit}
At low densities, i.e. for
large average inter-nucleon spacing, the interaction energy $W$
in Eq.~(\ref{Vmax}) is much smaller than the kinetic terms from
the free Hamiltonian. In case of the long wavelength limit,
Eq.~(\ref{sum1b}), we can therefore treat $H_T$ as a small
perturbation and write to lowest non-trivial order in $H_T$:
\begin{equation}
\langle H_T\rangle={2\over Z}\sum_n\exp\left(-E_n^0/T\right)
{\rm Re}\left[_1\!\left\langle n\right\vert H_T\left\vert
n\right\rangle_0\right]\,.\label{H31}
\end{equation}
Here, $E_n^0$, $\left\vert n\right\rangle_0$ are the eigenvalues and
eigenstates of the free Hamiltonian $H_0$, respectively,
$\left\vert n\right\rangle_1$ are the eigenvectors of $H_0+H_T$
to first order
in $H_T$, and $Z=\sum_n\exp\left(-E_n^0/T\right)$ is the normalization
factor. Assuming nondegenerate eigenstates for simplicity and
applying standard first order perturbation theory for
$\left\vert n\right\rangle_1$ we can express everything in terms
of zeroth order quantities.
Dropping the index $0$ from now on, Eq.~(\ref{H31}) reduces to
the negative definite expression
\begin{equation}
\langle H_T\rangle={1\over Z}\sum_{n\neq m}
{e^{-E_n/T}-e^{-E_m/T}\over
E_n-E_m}\left\vert(H_T)_{mn}\right\vert^2\,,\label{H32}
\end{equation}
where $(H_T)_{mn}=_0\!\left\langle m\right\vert H_T\left\vert
n\right\rangle_0$. This matrix element is expected to vary in
$E_m-E_n$ over a scale $\ga3m_\pi^2/M\simeq50\,{\rm MeV}$ where
$m_\pi$ is a typical momentum scale in the $NN$ interaction
potential. Therefore, for
$T\la50\,{\rm MeV}$ the thermal factor in Eq.~(\ref{H32})
can be approximated by $\delta(E_m-E_n)$. Converting the sum
over $m$ into an integral over $E_m$ Fermi's golden rule finally
gives $W\simeq-\langle H_T\rangle/N_b=\Gamma_\sigma/(2\pi)$. Here,
$\Gamma_\sigma$ is the average perturbative $NN$ scattering
rate mediated by $H_T$ which is a measure for
the spin fluctuation rate. The spins fluctuate on a time scale
given by the inverse energy scale of the tensor force
which causes the spin fluctuations. At high densities we use
$\Gamma_\sigma\equiv-2\pi\langle H_T\rangle/N_b$ as an effective
spin flip rate.
\section{Saturation of Spin Fluctuation Rates}
We now use the sum rules Eqs.~(\ref{sum1b}) and (\ref{sum2}) to
determine the qualitative form of $S_\sigma(\omega,{\bf k})$ in the long
wavelength limit. To this end let us introduce the dimensionless
quantity $\tilde{S}_\sigma(x)\equiv TS_\sigma(xT)$ with
$x=\omega/T$ as in
Ref.~\cite{Keil}. Due to the principle of detailed balance,
$S_\sigma(\omega,{\bf k})=S_\sigma(-\omega,-{\bf
k})e^{\omega/T}$, it is sufficient
to specify $\tilde{S}_\sigma(x)$ for $x>0$ only.
Introducing the dimensionless effective spin flip rate
$\gamma_\sigma=\Gamma_\sigma/T$, we can write the sum rule
Eq.~(\ref{sum1b}) as
\begin{equation}
\int_0^{+\infty}{dx\over2\pi}x\tilde{S}_\sigma(x)\left(1-e^{-x}\right)
={2\gamma_\sigma\over\pi}\simeq{4W\over T}\,,\label{sum1d}
\end{equation}
where in a newly born neutron star
$\gamma_\sigma$ does not increase beyond a few.
Furthermore, since in a hot SN core the thermal energies are
expected to be considerably higher than the interaction energy
$W$, within a first approximation we can neglect spin
correlations in the second sum rule Eq.~(\ref{sum2}) and write
\begin{equation}
\int_0^{+\infty}{dx\over2\pi}\tilde{S}_\sigma(x)\left(1+e^{-x}\right)
\simeq1\,.\label{sum2a}
\end{equation}
For the following discussion we consider the general case of an
ensemble of neutrons and protons with fractional number
densities $Y_n$ and $Y_p$. Introducing the isospin operators
${\hbox{\boldmath $\tau$}}_i$ for nucleon $i$,
${\hbox{\boldmath $\sigma$}}_i$
in the definition of $S_\sigma$ [see Eqs.~(\ref{sdef}) and (\ref{sk})]
has to be multiplied by $\left[1+(\tau_i)_3\right]C_{A,p}/2
+\left[1-(\tau_i)_3\right]C_{A,n}/2$. Here, $C_{A,p}$
and $C_{A,n}$ are the relevant proton and neutron axial-vector
charges. Moreover, there will be additional terms
proportional to ${\hbox{\boldmath $\tau$}}_i
\cdot{\hbox{\boldmath $\tau$}}_j$ in the interaction potential
Eq.~(\ref{Vint}). However, this leaves our discussion
qualitatively unchanged since the additional isospin operators appearing
under the expectation values only lead to additional factors of order
unity. If correlations among different nucleons are absent the
r.h.s. of the sum rules Eqs.~(\ref{sum1d}) and (\ref{sum2a}) get
multiplied by
$(Y_pC_{A,p}^2+Y_nC_{A,n}^2)/(C_{A,p}^2+C_{A,n}^2)$.
Parametrizing the high $\omega$ behavior of $S_\sigma$ by
$S_\sigma(\omega)\propto\omega^{-n}$, classical collisions would
lead to $n=2$. On the quantum mechanical level the deviation of
$S_\sigma(\omega)$ from $2\pi\delta(\omega)$ is to lowest order
in the strong interactions given by nucleon bremsstrahlung.
Using a dipole like OPE potential
without a hard core cutoff yields $n=5/2$ and $n=3/2$ in the case of
one and two nucleon species,
respectively~\cite{R2,FM,Iwamoto2,Brinkmann}. The non-existence
of the f sum Eq.~(\ref{sum1d}) in the latter
case stems from the unphysical $r^{-3}$ divergence
of this potential at $r=0$. Except for s waves this
divergence indeed leads to an infinite $\langle H_T\rangle$. If one
regularizes the potential by a hard core
repulsion f sum integrability is restored.
This motivates the following representative ansatz:
\begin{equation}
\tilde{S}_\sigma(x)={a\over x^{5/2}+b}\quad\hbox{for}\;x>0
\,,\label{St}
\end{equation}
where $a$ and $b$ are positive constants. The
sum rule Eq.~(\ref{sum1d}) is sensitive to the high energy
behavior and therefore mainly to $a$. In contrast, the sum rule
Eq.~(\ref{sum2a}) probes the ``infrared'' regime which is
sensitive to $b$. Eq.~(\ref{St}) is of the form expected from
nucleon bremsstrahlung where $b$ accounts for the LPM effect.
We can now pick
a number for $a$, determine the corresponding value of
$b$ numerically from Eq.~(\ref{sum2a}) and compute
the f sum Eq.~(\ref{sum1d}).
The result is plotted in Fig.~1 as a function of $a$.
Most importantly, from the expected density and temperature
dependence of $W$ we expect the f sum to increase monotonically
towards the SN core before saturating at a value of order unity.
As a consequence, the thermally averaged axial-vector neutrino
scattering cross section $\langle\sigma_A\rangle$ which
dominates the neutrino opacity should roughly scale
as $T^2$ being density independent as naively expected (see
Fig.~1). Furthermore, the axion emission rate from
Eq.~(\ref{Qa}) approximately scales as $n_b\Gamma_\sigma T^3$.
The lowest order axion emissivities should therefore be
multiplied by $\Gamma_\sigma/\Gamma_\sigma^\prime$ whenever this
ratio is smaller than 1. Here, $\Gamma_\sigma^\prime$ is the
lowest order spin flip rate extrapolated from the dilute medium
limit. For example, $\Gamma_\sigma^\prime\simeq32\,{\rm
MeV}\rho_{14}T_{10}^{1/2}$ for the standard OPE
calculations~\cite{FM,Iwamoto2,Brinkmann}, where $\rho_{14}$ is
the mass density in $10^{14}\,{\rm gcm}^{-3}$ and
$T_{10}=T/10\,{\rm MeV}$. A turn over in
$\langle\sigma_A\rangle/T^2$ and $Q_a/(n_bT^3)$ typically only
occurs at $\gamma_\sigma\ga10$ and is the less
pronounced the stronger $S_\sigma(\omega)$ falls off at large
$\omega$. The absence of a decrease of these quantities at high
density is therefore rather independent of
uncertainties in the exact saturation value for
$\gamma_\sigma$.
\section{Summary}
Neutrino opacities and axion
emissivities are governed mainly by the SSF. We have derived a
new sum rule for the SSF
which corresponds to the f sum rule for the density structure
function but depends on the nucleon spin flip
interactions. Our treatment so far
assumes absence of possible pion and kaon condensates. Employing
an infrared regularized bremsstrahlung spectrum for the
functional form of the SSF we have shown that the effective spin
fluctuation rate $\Gamma_\sigma$ must saturate somewhere
below $\simeq150\,{\rm MeV}$ which is within factors of a
few of SN core temperatures.
Neutrino scattering cross sections should therefore
exhibit the naive $T^2$ scaling whereas axion emissivities
should increase somewhat slower than the lowest order rates at
high densities. There is no turnover of weak
interaction rates towards the SN core. These results
have an important impact on SN cooling simulations and their
application to the derivation of axion mass bounds. They are also
relevant for the rates for URCA processes and emission of right
handed neutrinos.
\section*{Acknowledgments}
I gratefully acknowledge Georg Raffelt for an extensive e-mail
correspondence on many aspects of this research and for providing me
with early versions of Ref.~\cite{Janka}. I also thank
Thomas Janka for discussions of various astrophysical aspects.
This work was supported by the DoE, NSF and NASA at the
University of Chicago, by the DoE and by NASA through grant
NAG5-2788 at Fermilab, and by the Alexander-von-Humboldt
Foundation. Furthermore, I wish to thank the Aspen Center for
Physics where part of this research has been conducted for
hospitality and financial support.
|
\subsection*{\\Abstract}
\end{center}
We study the behavior of two diferent models at finite
temperature in a $D$-dimensional spacetime. The first
one is the $\lambda\varphi^{4}$
model and the second one
is the Gross-Neveu model.
Using the one-loop approximation we show that
in the $\lambda\varphi^{4}$ model the thermal
mass increase with the temperature while the thermal
coupling constant decrese with the temperature.
Using this facts we establish that in
the $(\lambda\varphi^{4})_{D=3}$ model
there is a temperature $\beta^{-1}_{\star}$ above which
the system can develop
a first order phase transition, where the
origin corresponds to a metastable vacuum.
In the massless Gross-Neveu model, we
demonstrate that for $D=3$ the thermal correction
to the coupling constant is zero. For $D\neq 3$ our
results are inconclusive.
\nopagebreak
Pacs numbers: 11.10.Ef, 11.10.Gh
\end{titlepage}
\newpage\baselineskip .37in
\section {Introduction}\
In the last years, there has been much interest in the
nature of the electroweak phase transition. The
high temperature effective potential in the
standard and in the
$(\lambda\varphi^{4})_{D=4}$ models
have been calculated by many authors, where the
contribution from multiloops diagrams has been
taking into account. Several
authors have pointed out the
importance to known whether in
$(\lambda\varphi^{4})_{D=4}$ model the phase transition is of
first or second order \cite{44}. Our
interest in these issues was stimulated by some
results of Ford and Svaiter
concerning the thermal dependence of the
mass and coupling constant in
$(\lambda\varphi^{4})_{D=4}$ model
defined in a non-simple connected spacetime \cite{1}.
In the aforementioned paper these authors studied a
neutral scalar field in a $D=4$
dimensional spacetime using the one-loop effective
potential. The
cases of trivial and non-trivial topology of the
spacelike sections and
finite temperature were discussed. The temperature
and topological dependent renormalized mass
and coupling
constant were derived
using the Speer and Bollini, Giambiagi
and Domingues analytic
regularization \cite{3} and a modified minimal
subtraction renormalization procedure \cite{40}. In
addition they have
also discussed
the possibility of vanishing the renormalized
coupling constant in this model,
as well as the limits of validity of the
one-loop approximation.
Some calculations studying such kind of problems
was given recently by Elizalde and Kirsten and also Villareal \cite{5}.
This last author improved the precedent results studying
the
two-loops corrections to the effective potential for
scalar fields defined in a spacetime
with non-trivial topology
of the spacelike sections.
The two goals of this paper are the following.
The first one is to extend
the discussion of
the massive self-interacting $\lambda\varphi^{4}$
model to an arbitrary
$D$-dimensional spacetime, assuming trivial
topology of the
spacelike sections and to analize temperature
effects in a model with
asymptotic freedom. The second one is to discuss
the existence of a first order
phase transition in the massive
$(\lambda\varphi^{4})_{D<4}$
model.
Besides Yang-Mills theories in $D=4$,
the other known
perturbative renormalizable
asymptotically free theories with fermions are the
Nambu-Jona-Lasinio and
the Gross-Neveu models
\cite{4}. In the latter, a $N$ component
fermion field with a quartic
self-interaction
is assumed. The model is perturbatively
renormalizable for $D=2$ and develops
asymptotic freedom.
Working in a generic
$D$-dimensional
spacetime, we first calculate the one-loop corrections to
the renormalized mass and coupling
constant
in the $\lambda\varphi^{4}$ model.
We obtained that the thermal mass increase and the
thermal coupling constant decrease with the temperature.
Still using
the one-loop aproximation, the thermal
correction to the renormalized coupling constant
in the Gross-Neveu model
is obtained. We demonstrate that in
the case $D=3$ the thermal
correction to the coupling constant
is zero. For $D\neq 3$ our results
are inconclusive.
In many papers studying second order
phase transition in the $\lambda\varphi^{4}$ model
the temperature dependence
of the coupling constant is neglected. This
approach is reasonable since the
variation of the mass with the temperature is the most
important fact for a critical phenomena.
In this case, it is sufficient to consider the
renormalized coupling constant
as constant and the thermal mass drives the
second order phase transition \cite{26}. In
this paper we will examine the
existence of a first order phase
transition in the $\lambda\varphi^{4}$
model taking into account the thermal dependence
of the coupling constant. Note that
we will not deal with the
system behavior in the neighborhood
of a second order phase transition since we assume that
the tree level mass squared $m^{2}$
is positive. This fact
prevents the one-loop approximation to break down
at low temperatures
since there is no infrared divergences associated with vanishing
masses. The result of our analysis can be summarized as
follows: for $D<4$, there is a temperature
$\beta^{-1}_{\star}$
where the effective coupling constant
vanishes. For temperatures $\beta^{-1}>\beta^{-1}_{\star}$, the
renormalized coupling constant becomes
negative and the system may suffer
a first order phase transition. The
effects of the radiative corrections
is toward the direction of breaking a symmetry.
Compare with
the electroweak first order phase transition \cite{7}.
We should note that at $\beta^{-1}=
\beta^{-1}_{\star}$ the
system is still in an interacting phase.
For $D<4$, there is
a temperature
where only the effective coupling constant
$(\lambda(\beta)=\lambda-\lambda^{2}f(\beta))$ vanishes.
All the higher
2n-points correlation functions
do not vanish, therefore the model is
not gaussian at the temperature
$\beta_{\star}^{-1}$. This is an important point that was
stressed by Weldon \cite{43}.
The study of the dependence
of the coupling
constant with the
temperature in QFT
is well known in the literature.
Many authors have studied
such dependence in the
$\lambda\varphi^{4}$ model and also in a abelian
model like QED \cite{41}. Instead
of using perturbative arguments, the use of the
renormalization group equations
allowed the investigation on the mass
and coupling constant thermal
dependence. Such program
was implemented by Fujimoto, Ideura,
Nakano and Yoneyama \cite{42}. These
authors obtained results similar to ours
in the $\lambda\varphi^{4}$ model. The behavior
of the mass and coupling constant
with the temperature are opposite, i.e. the
renormalized mass increases if the
system temperature increase as where the
coupling constant
decreases. If we assume that the
one-loop approximation provides
trustable results we have the
following situation: for
temperatures
above $\beta^{-1}_{\star}$ the
renormalized coupling constant
becomes negative. This
behavior of the effective coupling
constant is related to the fact
that the model is non-asymptotically free.
The growth of the renormalized
coupling constant at large momenta is
translated in our case to the
temperature growth (in modulus) of
this quantity. In $D=3$ for temperatures
$\beta^{-1}>\beta^{-1}_{\star}$ the system
develop
a first order phase transition
where the origin is a metastable vacuum.
In this paper we address only the
one-loop approximation. It is not
unreasonable to believe that our
conclusions in the $\lambda\varphi^{4}$
model may be limited to this approximation. In fact, the
behavior of the thermal correction to the
coupling constant changes in the
two-loops approximation. It was been
shown by Funakubo and Sakamoto \cite{41}
that only for {\it low}
temperatures the behavior of the
thermal coupling constant remains
the same as the obtained in the
one-loop approximation. For
high temperatures ($\beta^{-1}>>m$) the behavior is
opposite i.e., the thermal correction
is positive. Nevertheless this fact does not exclude the
possibility of a first order
phase transition at low temperatures in
$(\lambda\varphi^{4})_{D=3}$.
A more detailed discussion will appear in
a forthcoming paper.
The paper is organized as follows.
In section II we sketch
the formalism of the effective potential.
In section III, the massive
self-interacting $\lambda\varphi^{4}$ model
is analised.
In section IV we repeat the calculations
in the Gross-Neveu model.
Conclusions are given in section V .
In this paper
we use $\frac{h}{2\pi}=c=1$.
\section{ The effective action and the effective potential at
zero temperature.}
In this chapter we will review briefly the basic features of the
effective potential associated with a
real massive self-interacting scalar field at
zero temperature.
Although the formalism of this
section may be found in standard texbooks, we recall here its
main results for
completeness. Let us suppose a real
massive scalar field $\varphi(x)$ with the usual
$\lambda\varphi^{4}(x)$ self-interaction, defined in a static
spacetime. Since the manifold is static, there
is a global timelike
Killing vector field orthogonal to the
spacelike sections.
Due to this fact,
energy and thermal equilibrium have a precise meaning. For
the sake of simplicity,
let us suppose that the manifold is flat. In the path
integral approach, the
basic object is the generating functional,
$$
Z[J]= < 0,out | 0,in >=
$$
\begin{equation}
\int{\cal D}[\varphi]\exp \{i[S[\varphi]+
\int d^{4}x J(x)\varphi(x)]\}
\end{equation}
where $ {\cal D}[\varphi]$ is the functional
measure and $ S[\varphi]$ is the classical
action associated with the scalar field. The
quantity $ Z[J] $ gives the
transition amplitude from the initial
vacuum $ |0, in > $ to the final vacuum
$ |0,out > $ in the presence of some
source $ J(x)$, which is zero outside
some interval $ [-T,T] $ and inside this interval
was switched adiabatically
on and off. Since we are
interested in the connected part of the time ordered products
of the fields,
we take the connected generating functional $W[J]$,
as usual. This quantity is
defined in terms of the vacuum persistent amplitude by
\begin{equation}
e^{i W[J]}= <0, out|0, in >.
\end{equation}
The connected $n$-point function $ G_{c}(x_{1},x_{2},..,x_{n})$
is defined
by
\begin{equation}
G_{c}(x_{1},x_{2},..,x_{n})=\frac{\delta^{n} W[J]}{\delta J(x_{1})...
\delta J(x_{n})}|_{J=0}.
\end{equation}
Expanding $ W[J]$ in a functional Taylor series,
the n-order coefficient of this
series will be the sum of all connected Feynman
diagrams with $n$ external
legs, i.e. the connected Green's functions defined by eq.(3). Then
\begin{equation}
W[J]=\sum^{\infty}_{n=0}\frac{1}{n!}\int
d^{4}x_{1}..d^{4}x_{n}~ G^{(n)}_{c}(x_{1},x_{2}..
..x_{n}) J(x_{1})J(x_{2})..J(x_{n}).
\end{equation}
The classical field $\varphi_{0}$ is given by
the normalized vacuum
expectation value of the field
\begin{equation}
\varphi_{0}(x)=\frac{\delta W}{\delta J(x)}=
\frac{<0, out|\varphi(x)|0, in >_{J}}{<0, out|0, in >_{J}},
\end{equation}
and the effective action $\Gamma[\varphi_{0}]$ is
obtained by performing a
functional Legendre transformation
\begin{equation}
\Gamma[\varphi_{0}]=W[J]-\int d^{4}x J(x)\varphi_{0}(x).
\end{equation}
Using the functional chain rule and the definition
of $\varphi_{0}$ given
by eq.(5) we have
\begin{equation}
\frac{\delta\Gamma[\varphi_{0}]}{\delta\varphi_{0}}=-J(x).
\end{equation}
Just as $W[J]$ generates the connected Green's
functions by means of a
functional
Taylor expansion, the effective action can be
represented as a functional
power series around the value $\varphi_{0}=0$,
where the coeficients are
just the proper $n$-point
functions $ \Gamma^{(n)}(x_{1},x_{2},..,x_{n})$ i.e.,
\begin{equation}
\Gamma[\varphi_{0}]=
\sum^{\infty}_{n=0}\frac{1}{n!}\int d^{4}x_{1}d^{4}x_{2}..
.d^{4}x_{n}~\Gamma^{(n)}(x_{1},x_{2},..,x_{n})
~\varphi_{0}(x_{1})\varphi_{0}(x
_{2})..\varphi_{0}(x_{n}).
\end{equation}
The coefficients of the above functional expansion
are the connected $1$
particle irreducible diagrams $(1PI)$. Actually,
$\Gamma^{(n)}(x_{1},x_{2},..
.,x_{n})$ is the sum of all $ 1PI$ Feynman diagrams
with $n$ external legs.
Writing the effective action in powers of momentum
(around the point where
all external momenta vanish) we have
\begin{equation}
\Gamma[\varphi_{0}]=\int d ^{4}x\biggl(-V(\varphi_{0})+
\frac{1}{2}(\partial_{\mu}
\varphi)^{2} Z[\varphi_{0}]~+~..\biggr).
\end{equation}
The term $ V(\varphi_{0})$ is called the effective
potential\cite{8}\cite{9} .To express
$ V(\varphi_{0})$ in terms of the $1PI$
Green's functions, let us write
$ \Gamma^{(n)}(x_{1},x_{2},..,x_{n})$ in
the momentum space:
\begin{equation}
\Gamma^{(n)}(x_{1},x_{2},..,x_{n})=
\frac{1}{(2\pi)^{n}}
\int d^{4}k_{1}d^{4}k_{2}..d^{4}k_{n} (2\pi)^{4}
\delta(k_{1}+ k_{2} +..k_{n})\ e^{i(k_{1}x_{1}+..k_{n}x_{n})}
\tilde \Gamma^{(n)}(
x_{1},x_{2},..,x_{n}).
\end{equation}
Assuming that the model is translationally
invariant, i.e. $\varphi_{0}$
is constant over the manifold, we have
\begin{equation}
\Gamma[\varphi_{0}]=\int d^{4}x \sum^{\infty}_{n=1}
\frac{1}{n!}\biggr(\tilde\Gamma^{(n)}(0,0,..
.)(\varphi_{0})^{n}+...\biggr).
\end{equation}
If we compare eq.(9) with eq.(11) we obtain that
\begin{equation}
V(\varphi_{0})= -\sum_{n}\frac{1}{n!}\tilde\Gamma^{(n)}
(0,0,..)(\varphi_{0})^{n},
\end{equation}
then $ \frac{d^{n}V}{d\varphi^{n}_{0}} $ is the
sum of the all $1PI$ diagrams
carring zero external momenta. Assuming that the
fields are in equilibrium
with a thermal reservoir at temperature
$\beta^{-1}$, in the Euclidean
time formalism, the effective potential
$V(\beta,\varphi_{0})$ can be
identified with the free energy density and
can be calculated by
imposing periodic (antiperiodic) boundary
conditions on the bosonic (fermionic) fields.
In the next section using the effective potential
we will perform
the one-loop renormalization
of the $\lambda\varphi^{4}$ assuming that the system is
in equilibrium with a
thermal reservoir at temperature $\beta^{-1}$.
Since we are interested to
make a paralel with the tricritical phenomena where in the
tree level approximation with
$V(\varphi)=m^{2}\varphi^{2}+
\lambda\varphi^{4}+\sigma\varphi^{6}$ predicts
the existence of a first order phase transition if
we allow the coefficient of
the quartic term to be negative, we will
evaluate the effective potential in a very unusual way.
Instead of summing the series obtaining a log expression, and
regularizing the model by introducing an
ultraviolet cut-off in the Euclidean momenta,
we prefer to use the principle
of analytic extension in each
term of the series. The advantage of this
method lies in the fact that
the dependence of mass and coupling constant
with the temperature appear
in a very straightforward way as well as the paralel with
the tricritical phenomena.
\section{ The one-loop effective potential in
the $\lambda\varphi^{4}$ model
at zero and finite temperature.}\
Let
us assume the following
Lagrange density associated with a massive
neutral scalar field:
\begin{equation}
{\cal L}= \frac{1}{2}(\partial_{\mu}
\varphi_{u})^{2}-\frac{1}{2}m^{2}_{0}~
\varphi_{u}^{2}-\frac{\lambda_{0}}{4!}\varphi^{4}_{u}\,
\end{equation}
where $\varphi_{u}(x)$ is the unrenormalized
field and $m_{0}$ and $\lambda_{0}$
are the bare mass and bare coupling constant
respectively. We may rewrite the
Lagrange density as the usual form where the
counterterms will appear explicity.
Defining the quantities
\begin{equation}
\varphi_{u}(x)= (1+\delta Z)^{\frac{1}{2}}\varphi(x)
\end{equation}
\begin{equation}
m^{2}_{0}=(m^{2}+\delta m^{2}) (1+\delta Z )^{-1}
\end{equation}
\begin{equation}
\lambda_{0}= (\lambda+\delta\lambda)(1+\delta Z)^{-2},
\end{equation}
and substituting eq.(14),(15) and (16) in eq.(13) we have
\begin{equation}
{\cal L}=\frac{1}{2}(\partial_{\mu}\varphi)^{2}-
\frac{1}{2}m^{2}\varphi^{2}-
\frac{\lambda}{4!}\varphi^{4}+\frac{1}{2}
\delta Z(\partial_{\mu}\varphi)^{2}-
\frac{1}{2}\delta m^{2}\varphi^{2}-
\frac{1}{4!}\delta\lambda\varphi^{4},
\end{equation}
where $\delta Z$, $\delta m^{2}$, and
$\delta\lambda$ are the wave function,
mass and coupling constant counterterms of the model.
After the Wick rotation, in the one-loop
aproximation, the effective potential is given by \cite{9}:
\begin{equation}
V(\varphi_{0})=V_{I}(\varphi_{0})+V_{II}(\varphi_{0})
\end{equation}
where,
\begin{equation}
V_{I}(\varphi_{0})= \frac{1}{2}m^{2}\varphi^{2}_{0}+
\frac{\lambda}{4!}
\varphi^{4}_{0}-\frac{1}{2}\delta m^{2}\varphi^{2}_{0}-
\frac{1}{4!}\delta\lambda\varphi^{4}_{0},
\end{equation}
and
\begin{equation}
V_{II}(\varphi_{0})=
\sum_{s=1}^{\infty}\frac{(-1)^{s+1}}{2s}\biggl(\frac{1}{2}
\lambda\varphi^{2}_{0}\biggr)
^{s}\int\frac{d^{D}q}{(2\pi)^{D}}
\frac{1}{(\omega^{2}+\vec{q}~^{2}+m^{2})^s}.
\end{equation}
Before continuing, we would like to
discuss one important point.
Performing analytic or dimensional
regularization, we must
introduce a mass parameter $\mu$, in
terms of which dimensional
analysis gives to the field a dimension $[\varphi]=\mu^{1/2(D-2)}$
and to the coupling constant a dimension
$[\lambda]=\mu^{4-D}$.
Mass has dimension of inverse length, i.e. $[\mu]=[m]=L^{-1}$, and
the effective potential (the energy
density per unit volume)
has dimension of $L^{-D}$.
It is not difficult to extend the
results given by eqs.(19) and
(20) to finite temperature
states. After a Wick rotation, the functional integral
runs over the fields that
satisfy periodic boundary conditions in Euclidean time.
The effective action
can be defined as in the zero temperature case by a
functional Legendre
transformation. Regularization and renormalization
procedures follow
the same steps as in the zero temperature case.
Although the counterterms
introduced at finite temperature are the same as
in the zero temperature
case, the finite part of the physical parameters
are temperature dependent. In this
situation, since the sign of the thermal
correction to the coupling constant
is negative, the possibility of vanishing
the renormalized coupling constant
appears.
To study temperature effects we perform
as usual the following
replacement in the Euclidean region:
\begin{equation}
\int \frac{d\omega}{2\pi}\rightarrow\frac{1}{\beta}\sum_{n}
\end{equation}
and
\begin{equation}
\omega\rightarrow\frac{2\pi n}{\beta}
\end{equation}
where $\omega_{n}=\frac{2\pi n}{\beta}$ are
the Matsubara frequencies.
Defining the dimensionless quantities:
\begin{equation}
c^{2}=\frac{m^{2}}{4\pi^{2}\mu^{2}}
\end{equation}
and
\begin{equation}
(\beta\mu)^{2}=a^{-1},
\end{equation}
the Born terms plus one-loop terms
contributing to the effective
potential give,
$$V(\beta,\varphi_{0})=V_{I}(\varphi_{0})+
V_{II}(\beta,\varphi_{0}) $$
where,
\begin{equation}
V_{I}(\beta,\varphi_{0})=\frac{1}{2}m^{2}\varphi^{2}_{0}
+\frac{\lambda}{4!}
\varphi^{4}_{0}-\frac{1}{2}
\delta m^{2}\varphi^{2}_{0}-\frac{1}{4!}
\delta\lambda\varphi^{4}_{0},
\end{equation}
and
\begin{equation}
V_{II}(\beta,\varphi_{0})=\frac{1}{\beta}
\sum^{\infty}_{s=1}\frac{(-1)^{s+1}}
{2s}\biggl(\frac{\lambda}{8\pi^{2}}\biggr)^{s}
\biggl(\frac{\varphi_{0}}{\mu}\biggr)^{2s}\int
\frac{d^{d}q}{(2\pi)^{d}}
A^{M^{2}}_{1}
(s,a).
\end{equation}
The function
\begin{equation}
A^{c^{2}}_{N}(s,a_{1},a_{2},..,a_{N})=
\sum^{\infty}_{n_{1},n_{2}..n_{N}=-\infty}
(a_{1}n_{1}^{2}+a_{2}n^{2}_{2}+...+a_{N}n^{2}_{N}+c^{2})^{-s}
\end{equation}
is the inhomogeneous Epstein zeta function\cite{10}, and finally
$$
M^{2}=\frac{1}{4\pi^{2}\mu^{2}}(\vec{q}~^{2})+c^{2}.
$$
Note that the mass parameter $\mu$ introduced in eqs.(23) and (24)
will be used from now on, since we must have dimensionless functions
when working with analytic extensions.
Let us define the modified inhomogeneous Epstein zeta function as
\begin{equation}
E^{c^{2}}_{N}(s,a_{1},a_{2},..a_{N})=
\sum^{\infty}_{n_{1},n_{2},..n_{N}=1}
(a_{1}n^{2}_{1}+..+a_{N}n^{2}_{N}+c^{2})^{-s}.
\end{equation}
Defining
the new coupling constant and a new vacuum
expectation value of the
field $\phi$ (dimensionless for $D=4$),
\begin{equation}
g=\frac{\lambda}{8\pi^{2}}
\end{equation}
\begin{equation}
\frac{\varphi_{0}}{\mu}=\phi
\end{equation}
\begin{equation}
k^{i}=\frac{q^{i}}{2\pi\mu}
\end{equation}
we rewrite eq.(26) withouth use the definition
of the inhomogeneous
Epstein zeta function as,
\begin{equation}
V_{II}(\beta,\phi)=
\mu^{D}\sqrt{a}\sum^{\infty}_{s=1}
\frac{(-1)^{s+1}}{2s}g^{s}\phi^{2s}\sum^{\infty}_
{n=-\infty}\int d^{d}k\frac{1}{(an^{2}+c^{2}+\vec{k}~^{2})^{s}}.
\end{equation}
To regularize the model we will use a mix between
dimensional and zeta
function analytic regularizations. Let us first use dimensional
regularization\cite{11}. Using the well known result,
\begin{equation}
\int\frac{d^{d}k}{(k^{2}+a^{2})^{s}}=
\frac{\pi^{\frac{d}{2}}}{\Gamma(s)}\Gamma(s-\frac{d}{2})
\frac{1}{a^{2s-d}},
\end{equation}
eq. (32) becomes
\begin{equation}
V_{II}(\beta,\phi)=
\mu^{D}\sqrt{a}\sum^{\infty}_{s=1}\frac{(-1)^{s+1}}{2s}
g^{s}\phi^{2s}\frac{\pi^{\frac{d}{2}}}
{\Gamma(s)}\Gamma(s-\frac{d}{2})
\sum^{\infty}_{n=-\infty}
\frac{1}{(an^{2}+c^{2})^{s-\frac{d}{2}}}.
\end{equation}
Defining,
\begin{equation}
f(D,s)=f(d+1,s)=\frac{(-1)^{s+1}}{2s}
\pi^{\frac{d}{2}}\Gamma(s-\frac{d}{2})
\frac{1}{\Gamma(s)}
\end{equation}
and substituting eqs.(27) and (35) in eq.(34) we obtain,
\begin{equation}
V_{II}(\beta,\phi)=\mu^{D}\sqrt{a}\sum^{\infty}_{s=1}f(D,s)g^{s}
\phi^{2s}A^{c^{2}}_{1}(s-\frac{d}{2},a).
\end{equation}
As we will soon see, the terms $s\leq\frac{D}{2}$ are
divergent and we
will regularize the one-loop effective potential using
the Principle
of the Analytic Extension. Let us assume that
each term in the series of
the one-loop effective potential $ V(\beta,\phi)$ is the
analytic extension of these terms, defining in the
beginning in an open connected set.
To render the discussion more general, let us discuss the
process of the analytic continuation of the
modified inhomogeneous Epstein
zeta function given by eq.(28). For $ Re(s) > \frac{N}{2}$,
the $E^{c^{2}}
_{N}(s,a_{1},a_{2},..a_{N}) $ converges and
represent an analytic function
of $ s$, so $Re(s) > \frac{N}{2} $ is the
largest possible domain of the
convergences of the series. This means that in eq.(36)
in the case $D=4$ only the terms
$s=1$ and $s=2$ are divergent. The term $s=1$
is the divergent
one-loop diagram
of the connected two-point function and it
contributes with a quadratic
divergence. The $s=2$ term is the divergent
one-loop diagram of the connected four point
function, and it contributes to the effective
potential with a logarithmic
divergence. Using a Mellin transform it is
possible to find the analytic
extension of the modified inhomogeneous Epstein zeta function.
After some calculations using Kirsten's results
\cite{12}, we have:
\begin{equation}
V_{II}(\beta,\phi)=\mu^{D}\sum^{\infty}_{s=1}
f(D,s)g^{s}\phi^{2s}\sqrt{\pi}
\biggl(\frac{m}{2\pi\mu}\biggr)^{D-2s}\frac{1}
{\Gamma(s-\frac{d}{2})}
\biggl(\Gamma(s-\frac{D}{2})+4\sum^{\infty}_{n=1}
\biggl(\frac{mn\beta}{2}\biggr)^{s-\frac{D}{2}}
K_{\frac{D}{2}-s}(mn\beta)\biggr)
\end{equation}
where $K_{\mu}(z)$ is the Kelvin function \cite{13}.
It is not difficult to show that:
\begin{equation}
V_{II}(\beta,\phi)=
\mu^{D}\sum^{\infty}_{s=1}g^{s}\phi^{2s}h(D,s)
\biggl(\frac{1}{2^{\frac{D}{2}-s+2}}
\Gamma(s-\frac{D}{2})
(\frac{m}{\mu})^{D-2s}+\sum^{\infty}_{n=1}\biggl(\frac{m}
{\mu^{2}\beta n}\biggr)^{\frac{D}{2}-s}
K_{\frac{D}{2}-s}(mn\beta)\biggr)
\end{equation}
where:
\begin{equation} h(D,s)=\frac{1}{2^{\frac{D}{2}-s-1}}\frac{1}
{\pi^{\frac{D}{2}-2s}}\frac{(-1)^{s+1}}{s}
\frac{1}{\Gamma(s)}.
\end{equation}
If we
suppose that $D=4$, the
model is perturbatively renormalizable
and an appropriate choice
of $\delta m^{2}$ and $\delta \lambda$ will render the
analytic extension of the terms of the
series in $s$ in the effective
potential analytic functions in the neighbourhood
of the poles $s=1$ and $s=2$ respectively.
The idea to extend the definition of
an analytic function to a larger domain
(analytic extension) and subtract poles
was exploited by Speer, Bollini and
others. In the method used by Bollini,
Giambiagi and Domingues, a complex parameter
$s$ was introduced as an expoent of the
denominator of the loop expressions
and the integrals are well defined analytic
functions of the parameters in
the region $Re(s)>s_{0}$ for some $s_{0}$.
Performing an analytic extension of
the expression for $Re(s)\leq s_{0}$, poles
will appear in the analytic
extension and the final expression becomes
finite after a renormalization procedure.
To find the exact form of the counterterms let us
use the renormalization conditions
\begin{equation}
\frac{\partial^{2}}{\partial\phi^{2}}
V(\beta,\phi)|_{\phi=0}=m^{2}\mu^{2}
\end{equation}
and
\begin{equation}
\frac{\partial^{4}}
{\partial\phi^{4}}V(\beta,\phi)|_{\phi=0}=\lambda\mu^{4}.
\end{equation}
Since the vacuum expectation value of the field has
been chosen to be constant,
there is no need for wave function renormalization.
Substituting
eqs.(25),(38) and (39) in eqs.(40) and (41) it
is possible to find the
exact form of the countertems in such a way that
they cancel the polar
parts of the analytic extension of the terms $s=1$ and $s=2$.
Note that we are using a "modified" minimal
subtraction renormalization scheme where the mass
and coupling constant
counterterms are poles at the physical values of $s$.
It is straighforward to
show that both $\delta m^{2}$ and $\delta\lambda$ are
temperature independent.
If a model at zero temperature is renormalizable with
some counterterms it is
also renormalizable at finite temperature
with the same counterterms.
This result was obtained in all orders of perturbation
theory by Kislinger
and Morley \cite{22}. In the
neighbourhood of the poles $s=1$ and $s=2$, the regular
part of the analytic extension of inhomogeneous Epstein
zeta function has two contributions: one which is
temperature
independent and that can be absorbed by the
counterterms and
another that is temperature dependent and cannot be
absorbed by
the counterterms. It is clear that the temperature
dependent mass
is proportional to the regular part of
the analytic extension of
the inhomogeneous Epstein zeta function in
the neighborhood of
the pole $s=1$. The same argument can be
applied to the renormalized
coupling constant. The thermal contribution
to the renormalized
coupling constant is proportional to the
analytic extension of
the inhomogeneous Epstein zeta function
in the neighborhood of
the pole $s=2$. The choice of the
renormalization point $\phi=0$
implies that only the regular part in the
neighborhood of the
pole $s=1$ will appear in
the renormalized mass. In the next
section (where massless
self-interacting fermion fields are
studied) we will show that all
the terms of the series of
the effective potential contribute
to the renormalized mass
and coupling constant and the sign
of the thermal coupling
constant cannnot be computed
for $D\neq 3$. From the above
discussion we can write
\begin{equation}
-\tilde\Gamma^{(2)}(p=0,\beta,\lambda,m)=
m^{2}(\beta)=m^{2}+\Delta m^{2}(\beta)
\end{equation}
and
\begin{equation}
-\tilde\Gamma^{(4)}(p=0,\beta,\lambda,m)=
\lambda(\beta)=\lambda+\Delta\lambda(\beta),
\end{equation}
where $m^{2}(\beta)$ and $\lambda(\beta)$ are respectively
the temperature
dependent renormalized mass squared and
coupling constant. It can be
directly shown that the thermal contribution to the
renormalized mass squared is given by:
\begin{equation}
\Delta m^{2}(\beta)-
\Delta m^{2}(\infty)=\frac{1}{8\pi^{2}}\lambda
\sum^{\infty}_{n=1}\frac{m}{\beta n}K_{1}(mn\beta).
\end{equation}
Using the asymptotic representation of
the Bessel function
$K_{n}(z)$ for small arguments
$$ K_{n}(z)\cong\frac{1}{2}\Gamma(n)
(\frac{z}{2})^{-n}~~,z\rightarrow 0 ~~ n=1,2,..,$$
we obtain that at high temperatures the
temperature dependent
mass squared is proportional
to $\lambda\beta^{-2}$ \cite{19}.
The result given by eq.(44) was
also obtained by Braden \cite{21} using
Schwinger's proper time method.
The same author also discussed the
two-loop effective potential and
the problem of overlapping divergences
where the possibility of temperature
dependent counterterms appears.
Nevertheless these divergences must cancel as it was stressed by
Kislinger and Morley \cite{22}.
Based uppon the same arguments
previously used, the thermal
contribution to the renormalized coupling constant is given by:
\begin{equation}
\Delta\lambda(\beta)-\Delta\lambda(\infty)=-\frac{3}{8\pi^{2}}
\lambda^{2}\sum^{\infty}_{n=1}K_{0}(mn\beta).
\end{equation}
The Bessel function $K_{0}(z)$ is
positive and decreases for $z>0$.
Therefore let us present an
interesting result: the renormalized
coupling constant attains its
maximum at zero temperature
$(\beta^{-1}=\infty)$ and
decreases monotonically as the
temperature increases. In other words,
the thermal contribution
to the renormalized coupling constant
$\Delta\lambda(\beta)-\Delta\lambda(\infty)$ is negative, and
increases in modulus with the temperature.
The same result was
obtained by Fujimoto, Ideura,
Nakano and Yoneyama using the
renormalization group equations
at finite temperature \cite{42}.
Once we are discussing
thermal effects, in the limit of zero temperature the thermal
contribution to the mass and coupling
constant must vanish
$(\tilde\Gamma^{(2)}(p=0,\beta=\infty,\lambda,m)=-m^{2}$ and
$(\tilde\Gamma^{(4)}(p=0,\beta=\infty,\lambda,m)=-\lambda$).This
can be easily seen from eqs.(44) and (45).
Since the thermal
contribution to the renormalized coupling
constant is negative
someone could enquiry: is it possible for
the renormalized coupling
constant to vanish? Once $\Delta\lambda(\beta)$
is $O(\lambda^{2})$
and we assume $D=4$, it is not possible
to implement
such a mechanism for finite temperatures.
For $D<4$ the renormalized coupling constant
is not necessarily a
small quantity and it can even become a
large quantity, due to its positive
dimension $4-D$ in terms of the mass parameter
$\mu$ (or using the
language of critical phenomena, due to
its positive dimension $4-D$ in terms
of the
scale $\frac{1}{a}$ where $a$ is
the lattice spacing). Therefore
we conclude that in the neighbourhood of $D=4$,
the renormalized coupling
constant $\lambda(\beta)$
could vanish only for very high temperatures.
As we consider smaller spacetime
dimensions the temperature where
$\lambda(\beta)$ vanishes becomes
lower and lower.
For instance, for
$D=3$ we expect to find a finite temperature
$\beta^{-1}_{\star}$
such that the renormalized coupling constant
vanishes.
We note
that there is no discontinuity
in the behavior between the cases $D=4$ and $D<4$ as we
will see later (see
eq.(49)). For $D<4$ the model becomes
superrenormalizable and
only a finite number
set of graphs need overall counterterms. In the
one-loop aproximation for $D=4$ there are only two divergent
graphs and for $D<4$ there is only
one. This result can be
easily obtained by investigating eq.(38). In this equation the
divergent part of the effective
potential is given
by $\Gamma(s-\frac{D}{2})$
and for $D<4$ only the $s=1$ pole
will appear. In other words, for $D<4$
there is only finite coupling constant
renormalization at the one-loop
aproximation. The graph $s=2$ gives a
finite and negative contribution
to the coupling constant. For $D\geq 4$ the renormalization of
the coupling constant is obligatory
(note the presence of the pole in
$s=2$). Going back to the $D$-dimensional case, the
renormalization conditions also are
given by eqs.(40) and (41).
Using the renormalization conditions
in eq.(38), we can find the regular
part of the analytic extension which
gives a finite contribution to the
renormalized mass squared
$\Delta m^{2}(D,m,\lambda,\beta)$ and coupling
constant $\Delta\lambda(D,m,\lambda,\beta)$ in a $D$-dimensional
flat spacetime. We will simplify
the notation writing $\Delta m^{2}(\beta)$
and $\Delta\lambda(\beta)$. For
even $D$ they are given respectively by:
\begin{equation}
\Delta m^{2}(\beta)=\frac{\mu^{D-2}
\lambda}{2(2\pi)^{D/2}}\biggl(\frac{(-1)^
{\frac{D}{2}-1}}{(\frac{D}{2}-1)!}
\psi(\frac{D}{2})(\frac{m}{\mu})^{D-2}+
\sum^{\infty}_{n=1}\biggl(\frac{m}
{\mu^{2}\beta n}\biggr)^{\frac{D}{2}-1}
K_{\frac{D}{2}-1}(mn\beta)\biggr)
\end{equation}
and
\begin{equation}
\Delta\lambda(\beta)=-\frac{3}{2}\frac{\mu^{D-4}\lambda^{2}}
{(2\pi)^{D/2}}\biggl(\frac{(-1)
^{\frac{D}{2}-2}}{(\frac{D}{2}-2)!}
\psi(\frac{D}{2}-1)(\frac{m}{\mu})^{D-4}+
\sum^{\infty}_{n=1}\biggl(\frac{m}
{\mu^{2}\beta n}\biggr)^
{\frac{D}{2}-2}K_{\frac{D}{2}-2}(mn\beta)\biggr)
\end{equation}
where $\psi(s)=\frac{d}{ds}ln\Gamma(s)$.
For odd $D$, the first
term between parentesis in eqs.(46)
and (47) must be
replaced by $\Gamma(1-\frac{D}{2})
(\frac{m}{\mu})^{D-2}$ and
$\Gamma(2-\frac{D}{2})(\frac{m}{\mu})^{D-4}$
respectively.
The first terms between parentesis
of eq.(46) and eq.(47) are
temperature independent therefore
it is possible to isolate the thermal
contribution to the renormalized
mass and coupling constant in a
generic $D$-dimensional spacetime
in the one-loop aproximation.
Using eq.(46) and eq.(47) we obtain the
following contribution to the thermal mass
and coupling constant respectively:
\begin{equation}
\Delta m^{2}(\beta)-\Delta m^{2}(\infty)=
\frac{\mu^{D-2}\lambda}{2(2\pi)^{D/2}}
\sum^{\infty}_{n=1}\biggl(\frac{m}{\mu^{2}\beta n}\biggr)^
{\frac{D}{2}-1}K_{\frac{D}{2}-1}(mn\beta)
\end{equation}
and
\begin{equation}
\Delta \lambda(\beta)-
\Delta\lambda(\infty)=-\frac{3}{2}\frac{\mu^{D-4}
\lambda^{2}}{(2\pi)^{D/2}}\sum^{\infty}_{n=1}
\biggl(\frac{m}{\mu^{2}\beta n}\biggr)^
{\frac{D}{2}-2}K_{\frac{D}{2}-2}(mn\beta).
\end{equation}
These are among the main results of the paper.
Since $\Delta\lambda(\beta)-
\Delta\lambda(\infty) <0$ we may
have a temperature $\beta_{\star}^{-1}$
where $\lambda(\beta)$ vanish for
$D<4$. Our result is different from the
Frohlich result \cite{15}
in which all all the Green's functions of the theory for
$D>4$ correspond to a free field i.e. the
model is gaussian at zero temperature above
four spacetime dimensions.
In our case, the higher $2n$-point functions are not zero
as was discussed by Weldon
\cite{43}.
Before discussing a existence of a first order phase transition,
we would like to point out that the investigation of the
$(\lambda\varphi^{4})_{D=4}$ model with a negative bare
coupling constant has recently been done by Langfeld et al,
where an analytic continuation of
the model with positive $\lambda$
to negative values was presented \cite{18}. Although several
authors claim that the renormalized
coupling constant of the $\lambda\varphi^{4}$
model must
be positive, a definitive supporting argument
is still lacking. Previous investigations
have been done by many authors \cite{23}.
We would like to stress that
the sign of the renormalized coupling
constant is not fixed by the
renormalization procedure in the $(\lambda\varphi^{4})_{D=4}$.
Gallavoti
and Rivasseau discussed examples with
positive bare coupling constant
where different cutoffs lead to
renormalized coupling constants with
different signs \cite{24}.
Going back to the discussion of a first order
phase transition, let us
define a dimensionless effective
potential $v=\frac{V}{\mu^{D}}$,
as:
\begin{eqnarray}
v(\beta,\phi) &=& \frac{1}{2}m^{2}\mu^{2-D}\phi^{2} + \frac{\lambda}{4(2\pi)^{D/2}}\sum^{\infty}_{n=1}\biggl
(\frac{m}{\mu^{2}\beta n}\biggr)^{\frac{D}{2}-1}K_{\frac{D}{2}-1}
(mn\beta)\phi^{2}\nonumber \\
& + & \frac{\lambda}{4!}\mu^{4-D}\phi^{4}
-\frac{1}{16}\frac{\lambda^{2}}{(2\pi)^{D/2}}
\sum^{\infty}_{n=1}
\biggl(\frac{m}{\mu^{2}\beta n}\biggr)^
{\frac{D}{2}-2}K_{\frac{D}{2}-2}(mn\beta)\phi^{4}\nonumber\\
&+&\ high\ order\ terms\ in\ s.
\end{eqnarray}
In the effective potential all the powers $\phi^{2s}$
of the field will appear as stated in eq.(38).
For instance, the term
corresponding to the
$2s-th$ power of the field is proportional to
\begin{equation}
\sum^{\infty}_{n=1}\biggl(\frac{m}{\mu^{2}\beta n}\biggr)^
{\frac{D}{2}-s}K_{\frac{D}{2}-s}(mn\beta)\phi^{2s}.
\end{equation}
The previous results can be used to demonstrate a
first order phase transition in the
$(\lambda\varphi^{4})_{D=3}$ model.
To simplify our discussion
let us assume that is
possible to truncate the series of the effective potential in
$s=3$. These does not imply the assumption
that high order powers of the
field gives vanishing contributions.
They are simply neglected
as compared to the leading terms,
since we are interested in the profile
of the effective potential
near the origin. The coefficient
of $\varphi^{6}$ is positive
(one requires this to ensure that the truncated
effective potential is bounded from below).
For the sake of simplicity,
let us also assume that the coefficient of
the $\varphi^{6}$ is constant and given by
$\sigma$ for both cases $D=3$ and
$D=4$.
In these
cases the leading contributions
to the effective potential are respectively:
\begin{eqnarray}
v(\beta,\phi) &=& \biggl(\frac{1}{2}m^{2}+
\frac{\lambda}{4(2\pi)^{\frac{3}{2}}}
\sum^{\infty}_{n=1}\biggl(\frac{m}{\mu^{2}\beta n}\biggr)^
{\frac{1}{2}}K_{\frac{1}{2}}(mn\beta)\biggr)
\phi^{2}\nonumber\\
&+&\biggl(\frac{\lambda}{4!}-
\frac{\lambda^{2}}{16(2\pi)^{\frac{3}{2}}}
\sum^{\infty}_{n=1}\biggl(\frac{m}{\mu^{2}\beta n}\biggr)^
{-\frac{1}{2}}K_{\frac{1}{2}}(mn\beta)\biggr)
\phi^{4}+\sigma\phi^{6},
\end{eqnarray}
and
\begin{eqnarray}
v(\beta,\phi)&=&\biggl(\frac{1}{2}m^{2}+
\frac{\lambda}{16\pi^{2}}
\sum^{\infty}_{n=1}\biggl(\frac{m}{\mu^{2}
\beta n}\biggr)K_{1}(mn\beta)\biggr)\phi^{2}\nonumber\\
&+&\biggl(\frac{\lambda}{4!}-\frac{\lambda^{2}}{64\pi^{2}}
\sum^{\infty}_{n=1}K_{0}(mn\beta)\biggr)
\phi^{4}+\sigma\phi^{6}.
\end{eqnarray}
From the above discussion, for $D<4$ we obtain
the following profile for the
effective potential in the neighborhood of the
origin.
Bellow the temperature $\beta^{-1}_{\star}$,
the dimensionless
effective potential has only one global minimum.
Heating the system
above the temperature $\beta^{-1}_{\star}$, the renormalized
coupling constant would become negative and the
system can develop
a first order phase transition since the
expectation value of the order parameter
changes discontinuously by temperature effects.
The situation is similar to the Coleman-Weinberg
mechanism for massless fields. The effects of
the quantum corrections is towards the
direction of breaking a symmetry. Note
the similarity with the tricritical phenomena
where in the tree level
$(V(\varphi)=m^{2}\varphi^{2}+\lambda\varphi^{4}+
\sigma\varphi^{6})$ the model
develop a first order phase transition if we allow the
coefficient of the quartic term to be negative
\cite{48}.
In a detailed study, using the ring-improved one-loop effective
potential, Arnold and Spinosa \cite{50} showed
that even for temperature independent coupling constant, the
$\lambda\varphi^{4}$ model can develop
at the first sight a first order phase transition.
Nevertheless, these authors verified that
the contribution of
higher loop corrections dominates over the
one-loop ring improved
contributions. By these reasons, in this
approximation they cannot distinguish
between a first or a second order phase transition.
As we discussed in the introduction,
the thermal correction to the
coupling constant if we include high order loops
in the effective potential is positive for
high temperatures.
Nevertheless for low temperatures
the effective renormalized coupling
constant may become negative. In this case we
still have a first order phase transition.
From the above discussion,
we have obtained the following
result: in the massive $\lambda\varphi^{4}$ for $D<4$
for temperatures above $\beta^{-1}_{\star}$ the
effective potential will develops a local minimum
at the origin (a false vacuum) and a global one outside
the origin. In this case
the initial
metastable phase may decay to
a stable one by nucleation of bubbles.
The temperature is the parameter that drives
the first order phase
transition. Evaluating the ring diagrams Carrington and
Takahashi independently obtained
in a pure scalar model at $D=4$ results
which are consonant with ours results in $D=3$ \cite{33}.
\section{ The one-loop effective potential in the massless
Gross-Neveu model at finite temperature.}\
Our purpose throughout this section is to examinate
the behavior of the renormalized coupling constant in a
model involving fermions with a quartic interaction.
In two-dimensional
spacetime $(D=2)$ the
model is renormalizable and ultraviolet asymptotically free.
We will consider an N-component fermion
field where the limit of
large N will be investigated. As it was
discussed in ref.(4), due to the quartic nature
of the interaction, it is possible to
introduce an ultralocal
auxiliar scalar field $\varphi$ which is
formally equal to
$g\overline{\psi}\psi$ where $\psi(x)$ is the fermionic field,
in order to present the effective potential of the model.
As we did in section II, we suppose that
the quantum field is in
thermal equilibrium with a reservoir at
temperature $\beta^{-1}$.
We will show that for $D=2$ and $D=4$ inthe one-loop
approximation the
sign of the thermal
correction to the renormalized coupling
constant cannot be
calculated. On the other hand, for $D=3$ in the
one-loop aproximation the thermal correction to the
renormalized coupling constant is zero.
The Lagrange density of the massless model is given by:
\begin{equation}
{\cal L}(\overline{\psi},\psi, \varphi)=i\overline{\psi}
\gamma^{\mu}\partial_{\mu}\psi-\frac{1}{2}\varphi^{2}-
g\varphi\overline{\psi}\psi.
\end{equation}
Defining $\varphi_{0}$ as the vacuum expectation value of
$\varphi$, i.e. $\varphi_{0}=<0|\varphi|0>=
<0|g\overline{\psi}\psi|0>$,
the leading terms in the effective potential
for large N are given by the
tree-level graphs plus all one-loop graphs,
\begin{equation}
V(\varphi)=\frac{1}{2}\varphi_{0}^{2}
-iN\sum^{\infty}_{s=1}\frac{1}{2s}(g\varphi_{0})^{2s}
\int\frac{d^{D}q}{(2\pi)^{D}}\frac{1}{k^{2s}}.
\end{equation}
After a Wick rotation we identify the effective
potential as the free energy of the system. At zero
temperature the model has a spontaneous breakdown of the
chiral symmmetry where the fermions acquire mass. The
symmetry is restored at finite temperature by a second order
phase transition \cite{2}. This result can
be obtained by summing the series in the effective potential.
Since we are interested only in the thermal behavior
of the mass and coupling constant instead of repeat the
well known calculations we will adopt a very unusual road,
similar to the previous chapter, by regularizing each
term of the series in the effective potential before summing up.
To introduce finite temperature effects we
assume that the Grassmannian
integration in the path integral goes over
anti-periodic configurations in
Euclidean time. In the effective potential
this is equivalent to the
replacement given by eq.(21) and
\begin{equation}
\omega\rightarrow\frac{2\pi}{\beta}(n+\frac{1}{2}).
\end{equation}
Using eq.(33) and defining $f(D,s)$ by:
\begin{equation}
p(D,s)=\frac{1}{2^{2s+1}}\frac{1}{\pi^{2s-
\frac{d}{2}}}\frac{(-1)
^{s}}{s}\frac{\Gamma(s-\frac{d}{2})}{\Gamma(s)},
\end{equation}
it is not difficult to show that
$V(\beta,\varphi_{0})$ is given by:
\begin{equation}
V(\beta,\varphi_{0})= \frac{1}{2}\varphi_{0}^{2}+
N\sum^{\infty}_{s=1}p(D,s)(g\varphi_{0})
^{2s}\beta^{2s-D}
\sum^{\infty}_{n=-\infty}\frac{1}{(n+\frac{1}{2})^{2s-d}}.
\end{equation}
Note that we are using
dimensional regularization in
eq.(55) and it is well known that
for massless fields this technique
requires modification in
order to deal with infrared
divergences \cite{29}. Since we
are regularizing only a
$d=D-1$ dimensional integral, this
procedure is equivalent
to inserting a mass into the $d$ dimensional integral.
In other words, the Matsubara frequency plays the role of a
"mass" in the integral, provided we exclude the limit
$\beta\rightarrow\infty$, which means that we must
restrict ourselves to non-zero temperature.
Again, as in eq.(30), we can define a new field $\phi=
\frac{\varphi_{0}}{\mu}$ (no confusion must be done between
the present auxiliar scalar field and the previous scalar field).
Using eq.(24) we obtain
\begin{equation}
V(\beta,\phi)=\frac{1}{2}\mu^{2}\phi^{2}+
N\mu^{D}\sum^{\infty}_{s=1}
p(D,s)a^{\frac{D}{2}-s}(g\phi)^{2s}\sum^{\infty}_{n=
-\infty}\frac{1}{(n+\frac{1}{2})^{2s-d}}.
\end{equation}
The Hurwitz zeta function is defined as
\begin{equation}
\zeta(z,q)=\sum^{\infty}_{n=0}\frac{1}{(n+q)^{z}}
\end{equation}
for $Re(z)>1$ and $q\neq{0,-1,...}.$
For $q=1$ we recover the usual Riemann zeta function.
Defining:
\begin{equation}
r(D,s)=p(D,s)\biggl(\zeta(2s-d,\frac{1}{2})+(-1)^{2s-d}
\zeta(2s-d,-\frac{1}{2})-\frac{1}{2^{d-2s}}\biggr)
\end{equation}
the effective potential can be written as:
\begin{equation}
V(\beta,\phi)=\frac{1}{2}\mu^{2}\phi^{2}+N\mu^{D}\sum^
{\infty}_{s=1}r(D,s)a^{\frac{D}{2}-s}(g\phi)^{2s}.
\end{equation}
The effective potential is still baddly defined and it will
be regularized by the principle of analytic extension. The
function $r(D,s)$ is valid in the begining in an open
connected set of points, i.e. for $Re(z)>1$. Since we are
considering even non perturbative
renormalizable models, let us study
the cases $D=2,3$ and $4$. We would like
to stress that even
for the non perturbative renormalizable models
it is possible to make qualitative predictions
and we will
regularize and renormalize the model in the standard way.
A strong argument in favor of the study
of the Gross-Neveu
model is that the non-renormalizability does not appear in
the leading $\frac{1}{N}$ approximation for $D=3$.
After the analytic continuation, the effective potential
requires a renormalization procedure in the points $s=1,2..$
The renormalization condition which will fix the form of the
counterterm of the pole $s=1$ is:
\begin{equation}
\frac{\partial^{2}V}{\partial\phi^{2}}|_{\phi=cte}=\mu^{2}
\end{equation}
Due to infrared divergences, we must follow
Coleman and Weinberg
\cite{9} and choose the renormalization point at non-zero $\phi$.
In order to evaluate the renormalized effective potential it is
necessary to use the Hermite formula of
the analytic extension
for the Hurwitz zeta function given by \cite {16}
\begin{equation}
\zeta(z,q)=\frac{1}{2q^{z}}+
\frac{q^{1-z}}{z-1}+2\int^{\infty}_{0}
(q^{2}+y^{2})^{\frac{-z}{2}}
\sin(z\arctan\frac{y}{q})\frac{1}{e^{2\pi y}-1}dy.
\end{equation}
It is not difficult to show that
the thermal contribution to
the renormalized coupling constant is,
\begin{equation}
\Delta g(\beta)=N\mu^{D-2}\sum^{\infty}_{s=1}
r(D,s)(2s)
(2s-1)g^{2s}(\beta\mu)^{2s-D},
\end{equation}
where it is understood that the polar terms
in the summation have
been subtracted remaining just the
regular part of the analytic continuation.
The situation is different from the
massive $\lambda\varphi^{4}$ model,
since we have the contribution of all
terms of the series in $s$ and the sign
of the thermal contribution to the
renormalized coupling constant
cannot be easily obtained. Nevertheless, for
sufficiently small $g$ the leading term is $O(g^{2})$. In this case,
for $D=3$ and using the fact
that $\zeta(0,q)=\frac{1}{2}-q$,
we obtain that $\Delta g=0$.
We found here that there is no
thermal correction to the coupling constant at least in the
one-loop approximation.
Note that $\Delta
g(\beta)$ is still not well behaved.
The terms $s>\frac{D}{2}$
are divergent in the low temperature limit (the use of dimensional
regularization in the begining of the
calculations leads to this situation).
For $s<\frac{D}{2}$, the high temperature
limit of the model is
problematic due to the well known fact
that ultraviolet divergences
are worst as the spacetime dimension increases.
\section{Conclusions}
In this paper we studied the renormalization
program assuming that scalar
or fermionic fields are in equilibrium with
a thermal reservoir at
temperature $\beta^{-1}$. We have attempted
to analize the consequences of
the fact that not only the renormalized mass, but also
the renormalized coupling constant
acquire thermal corrections.
It is well known that if we have a one spatial
dimension compactified system at a finite temperature, which has a
spontaneous symmetry breaking there are two different
ways to restore the symmetry.
Since the compactification of one spatial
dimension gives us the well
known mechanism of topological generation
of mass, it is possible
to restore the symmetry by thermal or
topological effect. There is a very
simple way to interpret the origin of
the thermal and topological mass
and coupling constant. The
effective potential is not well defined.
Using the Principle of the Analytic
Extension, we regularize the model and
the introduction of counterterms
remove the principal part of the analytic extension,
and the model becomes
finite. Meanwhile, in the neighbourhood
of the poles, the regular part of the
analytic extension does not vanish.
These temperature
dependent regular part
around the poles $s=1$ and $s=2$ (for $D=4$) are identified with
the thermal correction to
the mass and coupling constant.
It
was proved that in the $\lambda\varphi^{4}$ model,
in the one-loop aproximation, the
thermal correction to the renormalized
mass is positive and
the thermal correction to the renormalized
coupling constant is negative.
In this case
the renormalized coupling constant
attains its maximum at zero temperature and
decreases monotonically as the temperature
increases. Since in $D=4$,
$\Delta\lambda(\beta)$ is $O(\lambda^{2})$
it is not possible to vanish the
renormalized coupling constant at a finite
temperature of the thermal
bath. For strong couplings
($D<4$) there is a finite
temperature where this can be achieved.
For temperatures $\beta^{-1}>\beta^{-1}_{\star}$
(negative coupling
constant) the system can develop a first order
phase transition, where
the origin is a false vacuum.
It is not all clear for us if at $D=4$ the system can
develop a first order phase transition. We are using the
following argument to disregard such possibility. As we
discussed in the introduction, in the two-loops approximation
at high temperatures the thermal correction to the
coupling constant is positive. The fact that in $D=4$ the
model has a small zero temperature
coupling constant eliminate the
first order phase transition in $D=4$.
We would like to emphasize that the massive $\lambda\varphi^{4}$
model does not belong to the same universality class of
the Ising model.
It is well known that it is possible to
compare the $\lambda\varphi^{4}$ model in continuous
$D$-dimensional Euclidean space with the Ising model. One
lattice formulation can be done and the
continuum limit of the model ($a\rightarrow 0$, where $a$ is the
lattice spacing) exist if the correlation length goes to infinite.
This fact implies that at the continuum limit of the
lattice model the system must suffer a second order phase
transition. In other words, close to the
critical temperature a D-dimensional Ising model has the same
correlation functions as those
for a field theory ($\lambda\varphi^{4}$ model) defined in a
D-dimensional Euclidean space near the critical
temperature. Since in the
paper we assume that the tree level mass squared is always positive
and we found that the thermal mass squared is also positive,
we are always far from the
critical temperature. By these reasons the system cannot fall into
the universality class of a Ising model.
The analysis of this paper suggest two
possible directions. First, we have to study
the decay of the metastable
ground state in the $(\lambda\varphi^{4})_{D<4}$ model
evaluating the nucleation rate per unit
volume in the system. The theory of bubbles nucleation at zero
and finite temperature was proposed and developed
by many authors \cite{47}. The basic result is that the
probability per unit volume per unit time of the metastable vacuum
to decay is given by $\Gamma=A e^{-S(\varphi)}$, where $S(\varphi)$ is
the Euclidean action of the "bounce" solution which describes tunneling
between a metastable and a true vacuum.
Another possible direction is to examinate if the
metastability of the system
(the false ground state) can be eliminated in a
more general scalar model. This former subject will
be presented soon in a
forthcomming paper\cite{36}. We conclude the paper
with some some questions which remain to be
answered.
(i) Is the existence of the first order phase transition in
$(\lambda\varphi^{4})_{D=3}$ an
artifact of our approximation? It will be interesting
to obtain a non-perturbative argument to demonstrate or disprove
this fact in a general way.
(i) Is the series given by eq.(69) Borel summable ?
It is well known that
the lack of Borel summability means that the system is unstable, since the
vacuum to vacuum amplitude develop and imaginary
part \cite{25}. It would be
interesting to investigate these questions.
\section{Acknowlegement}
We would like to thanks
Prof.L.H.Ford, Prof.A.Grib and Prof.C.A.Carvalho for useful
comments and criticisms and for valuable discussions. We are also
greateful to Dr.T.Vachaspati and Prof.A.Vilenkin
for many helpful discussions. N.F.Svaiter
would like to thanks the
hospitality of the Institute of Cosmology, Tufts University,
where part of this
work was carried out. This paper was
suported by Conselho Nacional de
Desenvolvimento Cientifico e Tecnologico do Brazil (CNPq).
|
\section{Introduction}
The quantum effects of the 2 dimensional (2d) gravitational theories are
recently measured numerically
in the computer simulation with high statistics.
In particular the data for the entropy exponent
(string susceptibility) in 2d
quantum gravity(QG) is the same as the known exact result within
a relative precision of $O(10^{-3})$. It is due to the developement of the
simulation technique in the dynamical triangulation\cite{ADF,D,KKM}
and the findings of new observables in QG such as MINBU distribution
\cite{JM,AJT,Th}.
The data analysis is done by a rather orthodox approach,i.e., the semiclassical
approximation. It has recently been applied to 2d $R^2$-gravity and
the simulation data of $<\intx\sqg R^2>$\ and its cross-over phenomenon
are successfully explained\cite{ITY}.
We list the merits of this approach.
\begin{enumerate}
\item
The semiclassical treatment
is, at present, the unique field-theoretical approach which can analyse
the mysterious region $(25\geq )c_m\geq 1$. The conformal field theory
gives a meaningful result only for some limitted regions of $c_m$.
The Matrix model is in the similar situation.
\item
Comparison with the ordinary quantization is transparent because
the ordinary renormalizable field theories ,such as QED and QCD,
are quantized essentially in the semiclassical way.
In particular,the renormalization properties of (2d) QG are expected to be
clarified in the semiclassical approch\cite{S1}.
\item
This approach can be used for the higher-dimensional QG such as
3d and 4d QG.
\end{enumerate}
The approach is perturbative, therefore choosing
the most appropriate vacuum under the global constraints (such as the area
constraint and the topology constraint)
is crucial in the proper evaluation. We explain it in Sect.3.
We add $R^2$-term to the ordinary 2d gravity
for the following reasons.
( We call the ordinary 2d gravity {\it Liouville gravity}
in contrast with {\it $R^2$-gravity} for the added one. )
\begin{enumerate}
\item
For the positive coupling,
the term plays the role of suppressing the high curvature and making
the surface smooth. For the negative one, the high curvature is energetically
favoured and making the surface rough. Therefore we can expect a richer
phase structure of the surface configuration.
\item
The term is higher-derivative ($\pl^4$), therefore it regularizes the
ultra-violet behaviour so good\cite{KPZ}. In fact the theory is
renormalizable\cite{S1}.
\item
The Einstein term ($R$-term) is topological in 2 dimension. It does not
have a local mode. The simplest interaction which is purely geometrical
and has local modes is $R^2$-term.
\item
In the lattice gravity, $R^2$-term is considered as one of natural irrelevent
terms in the continuous limit\cite{BK}.
\end{enumerate}
The $<\intx\sqg R^2>$\ simulation data
for $R^2$-gravity was presented by
\cite{TY} and the cross-over phenomenon was clearly found.
We present here MINBU
distribution data.
\section{Lattice Simulation of 2D Quantum R$^2$-Gravity
and MINBU Distribution}
The distribution of baby universe (BU) is one of important
observables in the lattice gravity\cite{JM,AJT,Th}. It was originally
introduced
to measure the entropy exponent (string susceptibility) efficiently.
Fig.1 shows the configuration of a BU with an area B (variable) from the mother
universe with an area A (fixed).
{\vs 6}
\begin{center}
Fig.1\q MINBU configuration
\end{center}
The 'neck' of Fig.1 is composed of three links which is
the minimum loop in the dynamically triangulated surface. The configuration
is called
the minimum neck baby universe (MINBU). MINBU distribution for the
Liouville-gravity and its matter-coupled case
were already measured\cite{AJT,Th,AT}.
First we explain briefly our lattice model of $R^2$-gravity.
The surface is regularized by the triangulation. The number of vertices
,where some links (edges of triangles) meet, is $N_0$. The number of
links at the i-th vertex ($i=1,2\cdots,N_0$) is $q_i$.
The number of triangles($N_2$) is related to $N_0$\ as
$N_2=2N_0-4$\ for the sphere topology.
The discretized model is then
described by
\begin{eqnarray}
&S_L=-\be_L\frac{4\pi^2}{3}\sum_{i=0}^{N_0}\frac{(6-q_i)^2}{q_i}
=-48\pi^2\be_L\sum_{i}\frac{1}{q_i}+\mbox{const}\com &\label{lat.1}
\end{eqnarray}
where $\be_L$\ is the R$^2$-coupling constant of the lattice model.
We do measurement for
$\be_L=0,50,100,200,300,-20,-50$\ .
We present the MINBU dstribution of $R^2$-gravity with no matter field
(pure R$^2$-gravity) in Fig.2 and 3 for
$\be_L\geq 0$\ and for $\be_L\leq 0$ respectively.
The total number of triangles
is $N_2=5000$.
For the detail see \cite{TY}.
{\vs 6}
\begin{center}
Fig.2\q MINBU distribution for $\be_L\geq 0$, Pure $R^2$-gravity.
\end{center}
{\vs 6}
\begin{center}
Fig.3\q MINBU distribution for $\be_L\leq 0$. Pure $R^2$-gravity.
\end{center}
As for positive $\be_L$\ (Fig.2), we see clearly the transition point $P_0$
, for each curve,
at which the distribution qualitatively changes. For the region $P=B/A\ > P_0$,
the birth probability
decreases as the size of BU increases.
For the region $P < P_0$,
the birth probability increases as the size
of BU increases.
The value of the transition point $P_0$\ depends on $\be$ \ and increases
as $\be$\ increases.
As for negative $\be$\ (Fig.3), the slope of the curve
tends to be sharp as $|\be|$\ increases at least for the region $P<P_1$.
The transition point $P_1$\ is not so clear as Fig.2.
In Sect 4.2 we interpret these data theoretically using the
semiclassical approach explained in Sect 3.
\section{Semiclassical Approach}
We analyse the simulation data by the semiclassical approach. The $R^2$-
gravity interacting with $c_m$-components scalar matter fields is
described by
\begin{eqnarray}
& S=\intx\sqg (\frac{1}{G} R-\be R^2-\mu
-\half\sum_{i=1}^{c_m}\pl_a\Phi_i\cdot
g^{ab}\cdot \pl_b\Phi_i)\com\q (\ a,b=1,2\ )\com & \label{3.1}
\end{eqnarray}
where
$G$\ is the gravitaional coupling constant, $\mu$\ is the cosmological
constant ,
$\be$\ is the coupling strength for $R^2$-term and $\Phi$\ is the $c_m$-
components scalar matter fields. The signature is Euclidean.
The partition function ,
under the fixed area condition\
$ A=\intx \sqg\ $ and with the conformal-flat gauge\
$g_{ab}=\ e^{\vp}\ \del_{ab}$\ , is written as
\cite{P},
\begin{eqnarray}
& {\bar Z}[A]=
\int\frac{\Dcal g\Dcal\Phi}{V_{GC}}\{exp\PLinv S\}~\del(\intx\sqg-A)
=exp\PLinv (\frac{8\pi(1-h)}{G}-\mu A)\times Z[A]\com & \nn\\
& Z[A]\equiv\int\Dcal\vp~ e^{+\frac{1}{\hbar}
S_0[\vp]}~\del(\intx ~e^\vp - A)\com & \label{3.2}\\
& S_0[\vp]=\intx\ (\frac{1}{2\ga}\vp\pl^2\vp
-\be~e^{-\vp}(\pl^2\vp)^2 +\frac{\xi}{2\ga}\pl_a(\vp\pl_a\vp)\ )\com
\q \frac{1}{\ga}=\frac{1}{48\pi}(26-c_m)\com & \label{3.3}
\end{eqnarray}
where $h$\ is the number of handles
\footnote{
The sign for the action is different from the usual convention as seen in
(\ref{3.2}).
}.
$V_{GC}$\ is the gauge volume due to the general coordinate invariance.
$\xi$\ is a free parameter. The total derivative term generally appears when
integrating out the anomaly equation
\ $\del S_{ind}[\vp]/\del\vp=\frac{1}{\ga}\pl^2\vp\ $.
This term turns out to be very important.
\footnote{
The uniqueness of this term, among all possible total derivatives, is shown
in \cite{ITY}.
}
We consider the manifold of a fixed topology of
the sphere ,$h=0$\ and the case $\ga>0\ (c_m<26)$.
\ $\hbar$\ is Planck constant.
\footnote{
In this section only,we explicitly write $\hbar$\ (Planck constant) in order
to show the perturbation structure clearly.
}
$Z[A]$\ is rewritten as, after the Laplace transformation and the inverse
Laplace one,
\begin{eqnarray}
& Z[A]=\int\frac{d\la}{\hbar}\int\Dcal\vp~exp\ \frac{1}{\hbar}[\ S_0[\vp]
-\la (\intx e^\vp - A)] &\nn\\
&=\int\frac{d\la}{\hbar}e^{\PLinv \la A}
\int\Dcal\vp~exp~\{\PLinv S_\la[\vp]\}\com &\nn\\
&S_\la[\vp]\equiv S_0[\vp]-\la\intx~e^\vp\ &\nn\\
&=\intx\ (\frac{1}{2\ga}\vp\pl^2\vp -\be~e^{-\vp}(\pl^2\vp)^2\
+\frac{\xi}{2\ga}\pl_a(\vp\pl_a\vp)\
-\la~e^\vp\ )\com &\label{3.4}
\end{eqnarray}
where the $\la$-integral should be carried out along an appropriate
contour parallel to the imaginary axis in the complex $\la$-plane.
Note that the $\del$-function
constraint in (\ref{3.2}) is substituted by the $\la$-integral.
The leading order configuration is given by the stationary minimum.
\begin{eqnarray}
\left. \frac{\del S_\la[\vp]}{\del\vp}\right|_{\vp_c}
=\left. \frac{1}{\ga}\pl^2\vp
+\be\{ e^{-\vp}(\pl^2\vp)^2-2\pl^2(e^{-\vp}\pl^2\vp)\}-\la e^\vp
\right|_{\vp_c}=0\com\nn\\
\left.\frac{d}{d\la}(\la A+S_\la[\vp_c])\right|_{\la_c}=0\com
\label{3.5}\\
Z[A]\approx \PLinv exp~\PLinv\{\la_cA+S_{\la_c}[\vp_c]\}\equiv
\PLinv exp~\PLinv \Ga^{eff}_c \pr\nn
\end{eqnarray}
Generally this approximation is valid for a large system.
In the present case, the system size is proportional to
$\frac{4\pi}{\ga}=\frac{26-c_m}{12}$. We expect the approximation
is valid except the region: $c_m\sim 26$.
The solution $\vp_c$\ and $\la_c$\ ,which describes the positive-constant
curvature solution and is continuous at $\be=0$,\ are given by\cite{ITY}
\begin{eqnarray}
\vp_c(r )=-ln~\{ \frac{\al_c}{8}(1+\frac{r^2}{A})^2\}\com\q
r^2=(x^1)^2+(x^2)^2\com \nn\\
\al_c=\frac{4\pi}{w}\{ w+1-\sqrt{w^2+1 -2\xi w} ~\}\com\q
w=16\pi\be'\ga\com\q \be'\equiv \frac{\be}{A}\com \label{3.6}\\
\ga\la_c A=\frac{w}{16\pi}(\al_c)^2-\al_c\com\nn
\end{eqnarray}
where $\xi$\ must satisfy
$-1\leq\ \xi\ \leq\ +1$\ for the realness of $\al_c$.
$(x^1,x^2)$\ are the flat (plane) coordinates.
The partition function at the classical level
is given by
\begin{eqnarray}
& \Ga^{eff}_c=~ln~Z[A]|_{\hbar^0}
=\la_c A+(1+\xi)\frac{4\pi}{\ga}~ln\frac{\al_c}{8}-\frac{\al_c}{\ga}w
+C(A)\com & \nn\\
& C(A)=\frac{8\pi (2+\xi)}{\ga}+\frac{8\pi\xi}{\ga}
\{~ln(L^2/A)-1~\}+O(A/L^2)\com &
\label{3.7}\\
& \frac{L^2}{A}\gg 1\com &\nn
\end{eqnarray}
where
$L$\ is the {\it infrared cut-off}
($r^2\leq L^2$)
introduced for the divergent volume integral of the total derivative term.
Note that $C(A)$\ does not depend on $\be$\ (or $w$) but on $c_m$\ (or $\ga$)
and $A$. Furthermore $C(A)$\ has an arbitrary constant of the form
$(8\pi\xi/\ga)\times\mbox{(const)}$\ due to the freedom of the choice of
the regularization parameter:\ $L\ra \mbox{(const)}'\times L$.
This arbitrary constant turns out to be important.
For the case $\be=0$\ , the theory is ordinary 2d gravity and we call it
Liouville gravity in contrast with $R^2$-gravity for $\be\not= 0$.
For the case $c_m=0$\ , the theory is called the pure gravity in contrast
with the matter-coupled gravity $c_m\not= 0$.
\section{Semiclassical Analysis of MINBU Distribution}
First we explain the free parameter $\xi$.
Recent analysis of the present theory at the (1-loop) quantum level has
revealed that it is conformal (the renormalization group beta functions=0)
for $w\geq 1$\ when we take $\xi=1$\ \cite{S1}.
Therefore the value $\xi=1$\ has some meaning purely within the theory.
The validity of this choice is also confirmed from a different approach,
that is,
the comparison of
the special case $\be$(or $w$)$=0$\ (Liouville gravity) of the present result
with the corresponding
result from the conformal field theory (KPZ result)\cite{KPZ}.
The asymptotic behaviour of $Z[A]|_{\hbar^0}$ \ at $w=0$\
is given, from (\ref{3.7}), as
\begin{eqnarray}
Z[A]|_{\hbar^0,w=0}=\left.e^{\Ga^{eff}_c}\right|_{w=0}
= exp\{ \frac{4\pi}{\ga}(3-\xi)
+(1+\xi)\frac{4\pi}{\ga}ln\frac{1+\xi}{2}+\frac{8\pi\xi}{\ga}ln\frac{L^2}{A}
\}\approx \nn\\
A^{-\frac{8\pi\xi}{\ga}}\times\mbox{const}=A^{-\frac{26-c_m}{6}\xi}
\times\mbox{const}\com\nn\\
\mbox{as}\ A\rightarrow +\infty\pr\label{cdep.1}
\end{eqnarray}
On the other hand, the KPZ result is
\begin{eqnarray}
Z^{KPZ}[A]\sim A^{\ga_s-3}\com\
\ga_s=\frac{1}{12}\{c_m-25-\sqrt{(25-c_m)(1-c_m)}\}+2\pr \label{cdep.2}
\end{eqnarray}
In order for our result to coincide with the KPZ result in the 'classical
limit' $c_m\rightarrow -\infty$\ :\
$Z^{KPZ}[A]\sim A^{+\frac{1}{6}c_m}$\ , we must take
\begin{eqnarray}
\xi=1\com \label{cdep.3}
\end{eqnarray}
in (\ref{cdep.1}).
In the following of this text we take this value.
\footnote{
In the numerical evaluation, we take $\xi=0.99$~ for the practical reason.
}
The asymptotic behaviour of
the present semiclassical result for the Liouville gravity
is, taking $\xi=1$\ in (\ref{cdep.1}),
\begin{eqnarray}
Z[A]\sim A^{-\frac{26-c_m}{6}}\times A^{-1}\com\q A\rightarrow +\infty
\com \label{cdep.4}
\end{eqnarray}
where the additional factor $A^{-1}$~ comes from the $\la$-integral in
the expression of $Z[A]$, (\ref{3.4})\cite{S2}.
Now we compare the KPZ result and the semiclassical result in the
normalized form.
\begin{eqnarray}
Z^{KPZ}_{norm}[A]\equiv \frac{Z^{KPZ}[A]}{Z^{KPZ}[A]|_{c_m=0}}
\sim A^{\ga_s(c_m)-\ga_s(c_m=0)}\com\nn\\
\ga_s(c_m)-\ga_s(c_m=0)=\frac{1}{12}\{c_m+5-\sqrt{(25-c_m)(1-c_m)}\}
\com \label{cdep.4b}\\
Z_{norm}[A]\equiv \frac{Z[A]}{Z[A]|_{c_m=0}}
\sim A^{+\frac{c_m}{6}} \pr \nn
\end{eqnarray}
We can numerically confirm that the semiclassical result, $\frac{c_m}{6}$,
and the KPZ result, $\ga_s(c_m)-\ga_s(c_m=0)$~, have very similar behaviour
for the region $c_m\leq 1$\cite{S2}.
Now we go back to the general value of $\be$.
The birth-probability
of the baby universe
with area $B (0 < B < A/2)$ from the mother universe with the total area A
is given by\cite{JM}
\begin{eqnarray}
{n_A(B)}=\frac
{3(A-B+a^2)(B+a^2)Z[B+a^2]Z[A-B+a^2]}
{A^2\times Z[A]} \nn\\
\approx\frac
{3(1-p)pZ[pA]Z[(1-p)A]}{Z[A]}\com \label{cdep.5}\\
\ln~(\frac{ {n_A(B)} }{3})
\approx\ln~(1-p)p+\ln~Z[pA]+\ln~Z[(1-p)A]-\ln~Z[A]\com\nn\\
p\equiv\frac{B}{A}\com\q 0<p<\half\pr \nn
\end{eqnarray}
We apply the result of $Z[A]$\ in Sect.3
to the above expressions.
\subsection{ $c_m$-dependence}
First we present the semiclassical prediction for Liouville gravity($\be=0$).
The result (\ref{3.7}) for the case $\be=0$\ gives ,taking $\xi=1$,
\begin{eqnarray}
\ga~ln~Z[rA]
=8\pi (\ln~\pi+1)+8\pi\ln(\frac{1}{r}\cdot \frac{L^2}{A})\pr
\label{cdep.6}
\end{eqnarray}
Then the MINBU distribution normalized by the pure garvity ($c_m=0$)
is obtained as
\begin{eqnarray}
\frac{n_A(B)}{n_A(B)|_{c_m=0}}=
\{p(1-p)\}^{\frac{c_m}{6}}\times
\exp~\{ \frac{c_m}{12}\times \Del\}\com\nn\\
\Del\equiv -2(\ln~\pi +1)-2\ln\frac{L^2}{A}\com
\label{cdep.7}
\end{eqnarray}
where $\Del$\ can be regarded as the free real parameter due to
the arbitrariness of
the infrared regularization parameter $L$.
We know from the result (\ref{cdep.7}) that the MINBU distribution lines
for different $c_m$'s cross at the single point $p=p^{*}$\ given by
\begin{eqnarray}
p^*(1-p^*)=\exp\{-\half~\Del\}\com\q p^*<\half\pr
\label{cdep.8}
\end{eqnarray}
Fig.4 shows three typical cases of $p^*$\ .
\vspace{2cm}
{\vs 6}
\begin{center}
Fig.4\q Three typical cases of the solution of (\ref{cdep.8}).
\end{center}
The choice of $\Del$\ is important to fit the theoretical curve (\ref{cdep.7})
with the data.
We show the behaviour of (\ref{cdep.7}) for the three
cases:\ 1)\ $\exp(-\half \Del)~\ll \fourth$\ ,Near Point O,Fig.5a\ ;\
2)\ $\exp(-\half \Del)~>\fourth$\ ,Above Point A ,Fig.5b\ ;\
3)\ $\exp(-\half \Del)~=\fourth -0$\ ,Near Point A ,Fig.5c.
{\vs 6}
\begin{center}
Fig.5a\q MINBU distribution for Liouville gravity, $\Del=8$
\end{center}
{\vs 6}
\begin{center}
Fig.5b\q MINBU distribution for Liouville gravity, $\Del=1$
\end{center}
{\vs 6}
\begin{center}
Fig.5c\q MINBU distribution for Liouville gravity, $\Del=3$
\end{center}
Fig.5a well fits with the known result of the computer
simulation\cite{AJT,AT}.
This result shows the importance of the infrared regularization.
\subsection{$\be$-dependence}
We consider the pure gravity($c_m=0$).
We plot MINBU dstribution, $ln~n_A(B)$,
as the function of $p\ (\ 0.001<p<0.1\ )$\
for various cases of $\be'=\be/A$\
($\xi=0.99$).
Fig.6a and 6b show that for $\be'>0$\ and $\be'<0$\ respectively.
{\vs 5}
\begin{center}
Fig.6a\q MINBU distribution for $\be'\geq 0$. $\xi=0.99,c_m=0$.
\end{center}
{\vs 5}
\begin{center}
Fig.6b\q MINBU distribution for $\be'\leq 0$. $\xi=0.99,c_m=0$.
\end{center}
The above results of Fig.6a and Fig.6b qualitatively coincide with
those of Fig.2 and Fig.3, respectively.
\vspace{1cm}
We list the asymptotic behaviour of $\ln~n_A(B)$\
for the general $\xi$\ and $c_m$\ in Table 1.
\vspace{0.5cm}
\begin{tabular}{|c|c|c|c|}
\hline
Phase & (C)\ $0<p\ll -w(\ltsim 1)$
& (B)\ $|w|\ll p$ & (A)\ $0<p\ll w(\ltsim 1)$ \\
\hline
$\al^-_p(pA)$ & $8\pi\{1+\frac{1-\xi}{2}\frac{p}{w}$
& $4\pi(1+\xi)\{1-\frac{1-\xi}{2}\frac{w}{p}$
& $\frac{4\pi(1+\xi)p}{w}\times $ \\
& $+O(\frac{p^2}{w^2})\}$ & $+O(\frac{w^2}{p^2})\}$
& $\{1+O(\frac{p}{w})\}$ \\
\hline
& $(1-\frac{8\pi\xi}{\ga})\ln~p$ & $(1-\frac{8\pi\xi}{\ga})\ln~p$
& $\{1-\frac{4\pi(1-\xi)}{\ga}\}\ln~p$ \\
$\ln~{n_A(B)}$ &$-\frac{4\pi}{\ga}\frac{w}{p}$
& $+O(\frac{w}{p})$
& $-\frac{4\pi(1+\xi)}{\ga}~\ln~w$ \\
&$+O(\frac{p}{w})$ & +SmallTerm
& $+O(\frac{p}{w}) $ \\
& +SmallTerm &
& +SmallTerm \\
\hline
\multicolumn{4}{c}{\q} \\
\multicolumn{4}{c}{Table 1\ \ Asymp. behaviour of
MINBU distribution,
(\ref{cdep.5}), }\\
\multicolumn{4}{c}{ for general $c_m$~ and $\xi$.
$R>0, w\equiv 16\pi\be'\ga, \ga=\frac{48\pi}{26-c_m}>0,p=\frac{B}{A},$}\\
\multicolumn{4}{c}{ $0<p\ll 1,\ |w|\ltsim 1,\ \mbox{SmallTerm}= \mbox{const}
+O(wp)+O(p).$}
\end{tabular}
\vspace{0.5cm}
We characterize each phase in Table 1 as follows.
\flushleft{(A)\ $0<p\ll w$:\ Smoothly Creased Surface
\footnote{
In \cite{ITY} we called it Free Creased Surface because this is the phase
where the free kinetic term ($R^2$-term) dominates.}
}
The smoothing term, $R^2$, dominates the main configuration and the surface
is smooth. The left part $P<P_0(w)$\ for each curve ($w$) in Fig.6a corresponds
to this phase.
The small BU is harder to be born because it needs high-curvature
locally. The large BU is energetically preferable to be born.
The area constraint is not effective in this phase.
The characteristic scale is $\be$.
\flushleft{(B)\ $|w|\ll p$:\ Fractal Surface}
The randomness dominates the configuration.
The size of BU is so enough large that the $R^2$-term is not effective.
The area constraint is neither effective. There is no characteristic scale.
The right part $P>P_0(w)$\ for each
curve ($w$) in Fig.6a and the right part $P>P_1(w)$\ for each curve ($w$) in
Fig.6b correspond to this phase.
The MINBU distribution is mainly determined
by the random distribution of the surface configuration\cite{BIPZ}.
\flushleft{(C)\ $0<p\ll -w$:\ Rough Surface
\footnote{
In \cite{ITY} we called it Strongly Tensed Perfect Sphere because the surface
tension is negatively large and the shape of the whole surface is near
a sphere. At the same time the surface tend to become
sharp-pointed because it increases the curvature.
We call the surface under this circumstace,simply, Rough Surface.}
}
Due to the large negative value of $R^2$-coupling, the configuration
with the large curvature is energetically preferable on the one hand,
it is strongly influenced by the area constraint on the other hand.
Therefore the large BU is much harder to be born than (B) because
it has a small curvature and a large area.
The small
BU is much easier to be born than (B) because it has a large curvature
and a small area. The left part $P<P_1(w)$\ for each curve ($w$) in Fig.6b
corresponds to this phase. The characteristic scale is the total area $A$.
\vspace{1cm}
\q We see the phase structure of Table 1 is the same as that of \cite{ITY}
by the substitution of $w$\ by $w/p$\ .
Although both simulations measure the same surface property,
the cross-over phenomenon,however, appears differently.
In \cite{ITY} the physical quantity
$<\intx\sqg R^2>$\ is taken to see the surface property.
The cross-over can be seen only by measuring for a range
of $w$\ and the transition point is given by a certain value
$|w^*|\approx 1$.
This is contrasting with the present case.
The cross-over can be seen
for any $w$. The transition is seen at the point $p^*$\ ,in the
MINBU distribution,
given by $|w|/p^*\approx 1$.
We understand as follows.
The MINBU distribution measures the surface
at many different 'scales' $B$, whereas the quantity $<\intx\sqg R^2>$\
measures the surface at a fixed 'scale'( $B_1$\ (or $p_1$) in the MINBU
terminology).
\subsection{ General Case}
We consider the general case of $c_m$\ and $\be$. This general case
is not yet measured by the Monte Carlo simulation. We present
the semiclassical prediction.
The analysis so far shows
the normalization
((\ref{cdep.4b}) and (\ref{cdep.7}))
and the choice of an arbitrary
constant due to the infrared regularization (\ref{cdep.7})
are important
for the quantitative adjustment.
Here, however, we are content with the qualitative behaviour.
We donot do the normalization and we ignore the $\ln~\frac{L^2}{A}$~term
in the evaluation of this subsection.
\flushleft{(1)\ $c_m$-dependence}
We stereographically show
MINBU distributions for the
range:\ $0.001\leq p\leq 0.2,\ -24\leq c_m\leq +24$\ ,
in Fig.7a($\be'=0$) , Fig.7b($\be'=+10^{-4}$) and
Fig.7c($\be'=-10^{-5}$).
{\vs 5}
\begin{center}
Fig.7a\q MINBU dstribution for $0.001\leq p\leq 0.2,\ -30\leq c_m\leq +24$\ .
$\be'=0,\xi=0.99$.
\end{center}
{\vs 5}
\begin{center}
Fig.7b\q MINBU dstribution for $0.001\leq p\leq 0.2,\ -30\leq c_m\leq +24$\ .
$\be'=+10^{-4},\xi=0.99$.
\end{center}
{\vs 5}
\begin{center}
Fig.7c\q MINBU dstribution for $0.001\leq p\leq 0.2,\ -30\leq c_m\leq +24$\ .
$\be'=-10^{-5},\xi=0.99$.
\end{center}
No 'ridge' appears in Fig.7a.
{}From this, we see matter fields
affect the surface dynamics homogeneously at all scales.
(This result is natural because the matter coupling constand $c_m$\ does
not have the scale dimension.) The slope along the $p$-axis continuously
decreases as $c_m$\ increases.
In Fig.7b,
a ridge runs from a low $p$\ to a high $p$\ as $c_m$\ increases.
In Fig.7c,a 'hollow' runs from a high $p$\
to a low $p$\ as $c_m$\ increases.
The ridge and the hollow correspond to the series of the cross-over points.
In both Fig.7b and Fig.7c, the cross-over
becomes dimmer as $c_m$\ increases and becomes sharper as $c_m$\ decreases.
\flushleft{(2)\ $\be$-dependence}
We stereographically show
MINBU distributions for the
range:\ $0.001\leq p\leq 0.2,\ -10^{-5}\leq \be'\leq +10^{-4}$\ ,
in Fig.8a($c_m=0$) , Fig.8b($c_m=+10$) and
Fig.8c($c_m=-10$).
{\vs 5}
\begin{center}
Fig.8a\q MINBU dstribution for $0.001\leq p\leq 0.2,
\ -10^{-5}\leq \be'\leq +10^{-4}$\ .
$c_m=0,\xi=0.99$.
\end{center}
{\vs 5}
\begin{center}
Fig.8b\q MINBU dstribution for $0.001\leq p\leq 0.2,
\ -10^{-5}\leq \be'\leq +10^{-4}$\ .
$c_m=+10,\xi=0.99$.
\end{center}
{\vs 5}
\begin{center}
Fig.8c\q MINBU dstribution for $0.001\leq p\leq 0.2,
\ -10^{-5}\leq \be'\leq +10^{-4}$\ .
$c_m=-10,\xi=0.99$.
\end{center}
The Fig.8a corresponds to the stereographic display of Fig.6a and 6b.
In each of Fig.8a-c, a ridge appears for $\be'>0$\ . For $\be'<0$\ ,
a tower appears instead of a ridge. For a large positive $c_m$\
( matter dominated region, $c_m=10$\ in Fig.8b) the undulation
of the MINBU dstribution surface
\footnote{Do not confuse it with the 2d manifold which
the present model of gravity represents.}
is small(the cross-over is dim), whereas it is large(the cross-over is sharp)
for a large negative $c_m$\
(matter anti-dominated region, $c_m=-10$\ in Fig.8c).
\section{Discussion and Conclusion}
In the (2d) QG,at present, there exists no simple way to find good
physical observables. They have been found by 'try and error'. MINBU is one
of good observables to measure the surface property. Quite recently
a new observable ,the 'electric resistivity' of the surface,
is proposed by \cite{KTY}.
By measuring the observable for
the matter-coupled Liouville gravity,
they observe a cross-over ,near $c_m=1$\ ,from the surface
where a complex-structure is well-defined to
the surface where it is not well-defined.
The analysis of
the new obserbable, from the standpoint of the present approach, is important.
\q There are some straightforward but important applications of
the present analysis
:\ 1)\ higher-genus case, 2)\ the case with
other higher-derivative terms such as $R^3$\ and $\na R\cdot \na R$\ ,
3)\ the quantum effect.
As for 2) ,references \cite{TY2} and \cite{Tsuda}
have already obtained the Monte Carlo data.
\q We have presented the numerical result of MINBU and its theoretical
explanation using the semiclassical approximation.
The surface properties are characterized.
It is confirmed that the present lowest approximation
is very efficient to analyse 2d quantum gravity, at least, qualitatively.
\q Finally we expect
other new observables will be found and many Monte Carlo
measurements will be done ,including 3 and 4 dimensional cases,
next a few years.
The interplay between the measurement by the
computer simulation and the theoretical interpretation will become important
more and more.
We believe this process
will lead to the right understanding of the (Euclidean) quantum gravity.
\begin{flushleft}
{\bf Acknowledgement}
\end{flushleft}
The authors thank N. Ishibashi and H. Kawai for comments and discussions
about the present work.
|
\section{INTRODUCTION}
In this paper we investigate matter-enhanced neutrino flavor
transformations, the MSW (Mikheyev-Smirnov-Wolfenstein) effect [1,2],
for the case of a matter density consisting of a smooth component and
a fluctuating part modeled by Gaussian colored noise. We will
consider the effects of fluctuation-induced neutrino flavor
decoherence during two post-core-bounce epochs of supernova evolution:
(1) the shock reheating epoch at TPB (time post bounce) $< 1$ s; and
(2) the hot bubble $r$-process nucleosynthesis epoch at TPB $\approx$
3--16 s.
The neutrino heating of the stalled shock has been examined by Fuller
et al. [3] and Wilson and Mayle [4] including neutrino flavor-mixing
effects. Flavor mixing between a light $\nu_e$ and $\nu_{\mu}$ or
$\nu_{\tau}$ with a cosmologically significant mass in the range of
10--100 eV can increase the supernova explosion energy by up to 60\%.
The increase in the explosion energy is a result of the greater
temperature of $\nu_\mu$ and $\nu_\tau$ compared to that of
$\nu_e$. Because of their charged current interactions with nucleons,
$\nu_e$ remain in thermal equilibrium with the matter to lower
densities and temperatures than $\nu_\mu$ and $\nu_\tau$. Thus,
although the neutrinos have approximately the same luminosities,
flavor transformation between $\nu_\mu$ or $\nu_\tau$ and $\nu_e$ on
their way to the stalled shock after thermal decoupling will increase
the average energy of $\nu_e$, resulting in more heat liberated behind
the shock than the case without neutrino flavor transformation.
This increase in the average energy of $\nu_e$ by the MSW
transformation also affects the possible $r$-process nucleosynthesis
of heavy elements in supernovae [5]. In the absence of flavor
transitions, $\bar\nu_e$ have a higher temperature than $\nu_e$. This
is due to the fact that the neutron-rich surface of the proto neutron
star presents a greater opacity to $\nu_e$ than $\bar\nu_e$.
Therefore, $\bar\nu_e$ decouple deeper in the proto neutron star than
$\nu_e$, and hence, are hotter. Due to their lack of charged current
interactions (at these energies) with nucleons, $\nu_\mu$ and
$\nu_\tau$ are even hotter than $\bar\nu_e$. As pointed out by Qian et
al. [5], an MSW transition can result in a proton production rate
which is greater than the neutron production rate at the proposed
$r$-process site. By requiring neutron-rich conditions at the
$r$-process site, Qian et al. were able to put limits on the vacuum
mass-squared difference and the vacuum mixing angle between $\nu_e$
and $\nu_\mu$ or $\nu_\tau$. Since $\nu_\mu$ and $\nu_\tau$ have
identical energy spectra, we will hereafter consider only MSW
transitions between $\nu_e$ and $\nu_\tau$ for convenience, although
all our results could equally well apply to MSW transitions between
$\nu_e$ and $\nu_\mu$.
A general, semi-quantitative approach to neutrino oscillations in
inhomogeneous matter was developed in Ref. [6]. Studies of matter
density fluctuations which are not random, but harmonic [7,8], or
occur as a jump-like change in the solar density [9], are available in
the literature. Inhomogeneities in the velocity field of material can
mimic the effects of density inhomogeneities on neutrino flavor
oscillations [8]. Velocity inhomogeneity effects become important if
the characteristic velocity is near the speed of light [8]. Noisy
mixing of matter could also mimic a fluctuating matter density.
A priori, fluctuations in the matter density may be well approximated
by random noise (which averages to zero) added to the average value of
the density. In Ref. [10], a differential equation for the averaged
survival probability was derived for the case in which the random
noise was taken to be a delta-correlated Gaussian distribution. It
was also shown that if the correlation length of the matter density
fluctuations was small compared to the neutrino oscillation length at
resonance, one obtained the same result as for the case of a
delta-correlated Gaussian. Here, we consider the more realistic case
of colored noise [11]. We find that the random fluctuations have the
largest effect on the flavor transition when the correlation length is
on the order of the neutrino oscillation length at resonance, and that
the fluctuations, on average, have the effect of suppressing the
flavor transition. This result is in agreement with more qualitative
arguments regarding flavor transition in an inhomogeneous distribution
of matter [6].
Suppressing the degree of flavor transition has the effect of lowering
the average energy of $\nu_e$ compared with the average energy they
would have had for an MSW transition in the absence of the noise. We
will show that neutrino flavor decoherence caused by random
fluctuations has little effect on the part of the neutrino mixing
parameter space $(\delta m^2,\
\sin^22\theta)$ previously excluded by considerations of MSW
neutrino flavor transformation in the hot bubble $r$-process region of
the supernova environment. However, we find that neutrino flavor
decoherence can have significant effects on the neutrino heating rate
and the electron fraction during the shock reheating epoch.
In Sec. II, we develop the equations for two-neutrino flavor
conversion in the presence of multiplicative colored noise. We
present a differential equation for the averaged transition
probability, which is valid when the correlation length is small
compared to the width of the resonance region. Section III presents
our results for the effects of random fluctuations on neutrino heating
of the shock and $r$-process nucleosynthesis in the neutrino-heated
supernova ejecta. In Sec. IV, we discuss our results and present our
conclusions.
\section{NEUTRINO FLAVOR EVOLUTION IN THE PRESENCE OF COLORED NOISE}
In this section, we discuss how an initially pure neutrino flavor
state propagating through a stochastic field of density fluctuations
can evolve into a mixed ensemble of neutrino flavors. We call this
process neutrino flavor decoherence (or flavor depolarization). Here
we consider two-neutrino mixing and we assume that there is the usual
unitary transformation between flavor eigenstates (e.g.,
$|\nu_e\rangle$ and $|\nu_\tau\rangle$) and mass eigenstates
($|\nu_1\rangle$ and $|\nu_2\rangle$):
\begin{equation}
|\nu_e\rangle = \cos\theta|\nu_1\rangle+\sin\theta|\nu_2\rangle,
\end{equation}
\begin{equation}
|\nu_\tau\rangle=-\sin\theta|\nu_1\rangle+\cos\theta|\nu_2\rangle,
\end{equation}
where $\theta$ is the vacuum mixing angle. Complete flavor
decoherence would occur if a neutrino emitted from the neutrino sphere
in an initially pure flavor state, for example $|\nu_e\rangle$,
becomes a mixed state of 50\% $|\nu_e\rangle$ and 50\%
$|\nu_\tau\rangle$ (both at the original neutrino energy) after
propagating through a region where density fluctuations exist.
Our goal in this section is to find which fluctuation characteristics
(e.g., root-mean-square amplitude, power spectrum, etc.) would be
required to cause significant deviations in neutrino flavor evolution
from that predicted by conventional MSW studies with a smooth density
run in the post-core-bounce supernova environment. We begin with a
discussion of the general flavor evolution problem and the way in
which fluctuations can be characterized.
The density matrix of two-neutrino flavor evolution obeys the equation
\begin{equation}
i{d\over dr}\hat {\rho} = [\hat H,\hat {\rho}],
\end{equation}
where
\begin{equation}
\hat {\rho} \equiv \pmatrix{
a_e(r) \cr
a_{\tau}(r)\cr}\otimes
(a^*_e(r),\ a^*_{\tau}(r)),
\end{equation}
with $a_e$ and $a_{\tau}$ the probability
amplitudes for the neutrino
to be $\nu_e$ and $\nu_\tau$, respectively, and where the Hamiltonian is
given by \cite{r1}
\begin{equation}
\hat H = \left({{-\delta m^2}\over 4E} \cos 2\theta + {1\over \sqrt{2}} G_F
(N_e(r)
+ N^r_e(r))\right){\sigma_z} + \left({{\delta m^2}\over 4E}
\sin 2\theta \right) {\sigma_x}.
\end{equation}
In Eq. (5), $\delta m^2$ is the vacuum neutrino mass-squared
difference, $E$ is the neutrino energy, $N_e$ and $N^r_e$ are the
averaged and randomly fluctuating parts of the electron number
density, respectively, and $\sigma_x$ and $\sigma_z$ are the Pauli
matrices.
For colored noise, we take the ensemble averages of the randomly
fluctuating part of the electron density to be given by
\begin{equation}
\langle N^r_e(r)\rangle = 0,
\end{equation}
\begin{equation}
\langle N^r_e(r)N^r_e(r^{\prime}) \rangle = {\beta}^2 \ N_e(r)
\ N_e(r^{\prime})
\ \exp(-|r-r^{\prime}|/\tau_c),
\end{equation}
with the averages of all odd products vanishing and all higher even
products given by all possible independent products of two-body
correlations (i.e., the fluctuations are Gaussian). For example, if we
define $f_{12 \cdots }=\langle N^r_e(r_1)N^r_e(r_2) \cdots \rangle$,
the average of a product of four would be $f_{1234}= f_{12}f_{34} +
f_{13}f_{24} + f_{14}f_{23}.$ In Eq. (7), $\beta$ is the ratio of the
root-mean-square fluctuation to the local density, and $\tau_c$ is the
correlation length.
Clearly, the detailed evolution of the position (time) dependent
amplitudes $a_e(r)$ and $a_\tau(r)$ through regions which include MSW
resonances (mass level crossings) is quite complicated in the presence
of fluctuations and so necessitates numerical treatment. However,
before we proceed to the description of the numerical calculation, it
is advantageous to define a few quantities which we shall later
employ. First, we introduce the resonance width as
\begin{equation}
\delta r=2H\tan2\theta,
\end{equation}
where the effective weak charge
density scale height is $H\equiv|d\ln N_e(r)/dr|^{-1}$ [1]. We define
$\Delta\equiv\delta m^2/2E$, where $\delta m^2$ is the difference of
the squares of the vacuum neutrino mass eigenvalues. With this
notation, resonance (neutrino mass level crossing) occurs at the
position where $\Delta\cos2\theta=\sqrt{2}G_FN_e$ is satisfied.
We now proceed to a detailed description of how we follow numerically
the time evolution of the neutrino density matrix [i.e.,
Eq. (3)]. Provided one has access to a random number generator which
generates Gaussian deviates, Eq. (3) can be integrated by the method
of Ref. [11] with the probabilities calculated and averaged. However,
such a process can be quite time consuming. For the case in which the
correlation length is small compared with the width of the resonance
region, a ``ladder'' approximation can be made allowing one to obtain
a differential equation for the averaged probabilities.
In order to obtain this approximation, one can transform $\hat
{\rho}$ to the interaction picture, express $\hat\rho_I$ as an
iterative expansion and explicitly perform the averaging to obtain
[10]
\begin{equation}
\langle\hat {\rho}_I(r)\rangle = {1\over 2}\left({1 + \sigma_z} \right)
+\sum_{n=1}^{\infty} (-1)^n \int dR(2n)
F(2n) \hat C (2n),
\end{equation}
where
\begin{equation}
\int dR(2n) \equiv \int_0^r dr_1 \int_0^{r_1} dr_2 \cdots \int_0^{r_{2n-1}}
dr_{2n},
\end{equation}
\begin{equation}
F(2n) \equiv f_{123 \cdots (2n-1) (2n)},
\end{equation}
\begin{equation}
\hat C (2n) \equiv [\hat M(r_1),[\hat M(r_2),[ \cdots, [\hat M(r_{2n}),
\sigma_z] \cdots],
\end{equation}
and $\hat M = (G_F/2\sqrt{2})\hat U_0^{\dagger} \sigma_z \hat U_0$,
with $\hat U_0$ being the propagator of the flavor evolution in the
absence of fluctuations. The averaged density matrix is then given by
$\langle\hat\rho\rangle=U_0\langle\hat\rho_I\rangle U_0^\dagger$.
{}From the physical fact that the fluctuations should not affect the
evolution far from resonance, one can restrict the integration limits
in Eq. (10) to the resonance region. Since the matrix elements of
$\hat C(2n)$ are always $\leq 1$, we evaluate the integrals of the
terms in $F(2n)$ and compare the relative size of these terms. When
$\tau_c \ll \delta r$, the leading term in the integral of $F(2n)$ is
of order $(\tau_c\delta r)^n$. This contribution comes from the term
in which the subscripts are in the order of the nested integrals
[i.e., ($f_{12}f_{34}
\cdots f_{(2n-1) (2n)})$]. Terms which are out of this order by $s$
interchanges are
of order $(\delta r)^{n-s} \tau_c^{n+s}$ when $\tau_c \ll \delta r$.
Note that because of the absolute value signs in Eq. (7),
$f_{ij} = f_{ji}$, so that $f_{13}f_{24}
\cdots f_{(2n-1) (2n)}$ is out of order by one interchange and
$f_{13}f_{45}f_{26}
\cdots f_{(2n-1) (2n)}$ is out of order by two interchanges.
Therefore, the integral of all terms in $F(2n)$
{\it not in} the order of the nested integrals will be of order
$\tau_c/\delta r$
or smaller than the integral of the term in $F(2n)$ which is {\it in} the order
of the nested integrals when $\tau_c \ll \delta r$. Therefore,
we approximate Eq. (9) by keeping only the largest term in
each $F(2n)$. One then
obtains an infinite series which is the iterative expansion of
\begin{eqnarray}
\langle\hat {\rho}_I(r)\rangle & = & {1\over 2}\left({1 + \sigma_z} \right)
- \int_0^r dr_1 \int_0^{r_1} dr_2\,
\beta^2 N_e(r_1) N_e(r_2) \exp[{-(r_1 - r_2)/\tau_c}] \nonumber\\
& \times & [\hat M(r_1), [\hat M(r_2),\langle\hat {\rho}_I(r_2)\rangle]],
\end{eqnarray}
where we have returned to the full integration limits since the
contribution from positions outside the resonance region is
small. Equation (13) could also have been obtained by making the
assumption $\langle N_e^r N_e^{r\prime}\hat {\rho}_I\rangle \sim
\langle N_e^r N_e^{r\prime}\rangle\langle\hat {\rho}_I\rangle$. Since
the maximal reduction in the transition probability occurs for cases
in which the correlation length is approximately the neutrino
oscillation length at resonance divided by $\pi$ (i.e., $\tau_c\sim
L_{\rm res}/\pi = 4E/\delta m^2 \sin 2\theta$), the above
approximation should be valid for adiabatic transitions for which the
oscillation length at resonance is much smaller than the width of the
resonance region.
As an example of the approximation, we calculate and present the
survival probability as a function of the correlation length for a
$\nu_\tau$ traveling through the supernova at TPB $\approx3$ s in
Fig. 1(a). The rms value of the noise is taken to be 1\% of the local
electron number density, and we choose the following parameters:
$\delta m^2 = 10^2$ eV$^2$, $\sin^2 2\theta = 10^{-3}$, and $E = 33$
MeV. The correlation length of the random fluctuations varies from
$\tau_c =$ (0.25--24)$\tau_c^0$, where $\tau_c^0 = L_{\rm res}/\pi =
({{\delta m^2}}\sin 2\theta/4E)^{-1} = 832.6$ cm. For these
parameters, the neutrino flavor evolution in the absence of noise is
highly adiabatic ($L_{\rm res}/ \delta r \sim 0.03$). In Fig. 1, the
smooth curve is the approximation and the jagged line is the
simulation of Eq. (9) utilizing numerical averaging. The survival
probability in the absence of fluctuations is not shown in Fig. 1(a),
being $\sim 0$. The agreement between the approximation and the
simulation is quite good. We also plot in Fig. 1(b) the $\nu_\tau$
survival probability as a function of the correlation length for the
same parameters as in Fig. 1(a), except with $\sin^2 2\theta =
10^{-5}$ and the corresponding $\tau_c^0 = 8326.0$ cm. In Fig. 1(b),
the horizontal line is the survival probability in the absence of
fluctuations. One observes that at small values of the correlation
length [$\tau_c \sim$ (0--5)$\tau_c^0$], the survival probability from
the simulation is larger than that calculated from the approximation,
although they are not greatly different.
The reason the approximation works as well as it does is that for both
$\tau_c \gg L_{\rm res}$ and $\tau_c \ll L_{\rm res}$, the
approximation and the numerically-averaged simulation approach the
probability in the absence of fluctuations. As discussed in Ref. [10],
the parameter $\gamma = 1/2 G_F^2 \langle(N_e^r)^2\rangle \tau_c
\delta r$ governs the size of the effect of the fluctuations when
$\tau_c \ll L_{\rm res}$. When $\gamma$ is large, the transition
probability is heavily suppressed, where as for $\gamma \ll 1$, the
fluctuations have little effect. Therefore, a small value for $\tau_c$
will result in only a small change in the probability. As $\tau_c$
goes to infinity,
\begin{eqnarray}
\lim_{\tau_c\to\infty}F(2n) & = & (2n-1)!! \beta^{2n}
\prod_{i=1}^{2n}
N_e(r_i) \nonumber \\
& = & {1\over{\sqrt{2\pi \beta^2}}} \int_{-\infty}^{\infty} dx
\exp[{-x^2/(2\beta^2)}] x^{2n} \prod_{i=1}^{2n} N_e(r_i),
\end{eqnarray}
and Eq. (9) can be summed and transformed back from the
interaction picture to give
\begin{equation}
\lim_{\tau_c\to\infty}\langle\hat \rho(r)\rangle =
{1\over{\sqrt{2\pi \beta^2}}} \int_{-\infty}^{\infty} dx
\exp[{-x^2/(2\beta^2)}]
\hat \rho(r,x),
\end{equation}
where $\hat \rho(r,x)$ is the density matrix calculated using Eq. (3)
for the electron density $(1 + x)N_e(r)$. Equation (15) provides a
simple physical picture for the averaged density matrix in the limit
of very large correlation lengths: one simply calculates the density
matrix for the electron density with a ``frozen'' fluctuation,
$(1+x)N_e(r)$, then averages over all such fluctuations with a
Gaussian weight. Large correlation lengths imply that fluctuations at
many different locations are coupled. Indeed, an averaging such as
given in Eq. (15) is common in other physical situations when many
coupled channels are present. For example, in multidimensional
dissipative quantum tunnelling, barrier transmission probability is
given by a similar formula when the number of channels gets very large
[12]. Note that to first order in $x$, the change
$N_e(r)\to(1+x)N_e(r)$ in the functional form of the density will
affect the survival probability via a change in the slope of the
density at resonance. For slowly changing slopes and small $\beta$,
the change in the survival probability should be proportional to
$\beta^2$, and therefore be quite small.
\section{EFFECTS ON THE $r$-PROCESS AND \hfill \break SHOCK REHEATING}
As demonstrated by Fig. 1(b), even a 1\% fluctuation in the matter
density (a value which is probably unrealistically large for the
rather quiescent TPB $>3$ s epoch) does little in reducing the
transition probability (or increasing the survival probability) for
non-adiabatic evolution. As shown in Ref. [5], the neutrino-heated
supernova ejecta must have a neutron excess ($Y_e<0.5$) in order for
any $r$-process nucleosynthesis to be produced. In Ref. [5], it was
shown that the electron fraction $Y_e$ is approximately given by
\begin{equation}
Y_e \approx {1 \over {1 + \lambda_{\bar{\nu}_e p}/
\lambda_{\nu_e n}}} \approx {1 \over {1 + \langle{E}_{\bar{\nu}_e}\rangle/
\langle{E}_{\nu_e}\rangle}},
\end{equation}
where $\lambda_{\bar{\nu}_e p}$ and $\lambda_{\nu_e n}$ are the
reaction rates for
\begin{eqnarray}
\bar{\nu}_e + p & \rightarrow & e^+ + n,\\
\nu_e + n & \rightarrow & e^- + p,
\end{eqnarray}
respectively, and $\langle E_{\nu_e}\rangle$ and
$\langle{E}_{\bar\nu_e}\rangle$ are the average energy for $\nu_e$ and
$\bar\nu_e$, respectively. Since at this epoch the average neutrino
energies are $\langle E_{\nu_e}\rangle\approx11$ MeV, $\langle
E_{\bar\nu_e}\rangle\approx16$ MeV, and $\langle
E_{\nu_\tau}\rangle\approx25$ MeV, a substantial conversion of
$\nu_\tau$ into $\nu_e$ can decrease the ratio
$\langle{E}_{\bar{\nu}_e}\rangle/ \langle{E}_{\nu_e}\rangle$ below one
and thereby make conditions impossible for the $r$-process. The
average $\nu_e$ energy after an MSW transition is approximately
$P(\nu_e \rightarrow \nu_e) \langle{E}_{\nu_e}\rangle + P(\nu_{\tau}
\rightarrow \nu_e)\langle{E}_{{\nu}_\tau}\rangle$. To obtain
neutron-rich conditions, one must have $P(\nu_e \rightarrow \nu_e) >
64\%$. It only requires about 30\% efficiency in flavor conversion to
drive the neutrino-heated supernova ejecta too proton rich for
$r$-process nucleosynthesis. Consider simple estimates for $Y_e$ using
Eq. (16) for four cases:
(1) No flavor mixing. In this case, we have $Y_e\approx1/(1+\langle
E_{\bar\nu_e}\rangle/ \langle
E_{\nu_e}\rangle)\approx1/(1+16/11)\approx0.41$, so the material is
neutron rich, in good agreement with detailed supernova model
calculations [5].
(2) Full flavor conversion. In this case, we have $\langle
E_{\nu_e}\rangle\rightleftharpoons\langle E_{\nu_\tau}\rangle$, so
that $Y_e\approx1/(1+\langle E_{\bar\nu_e}\rangle/ \langle
E_{\nu_\tau}\rangle)\approx1/(1+16/25)\approx0.61,$ and the material
is very proton rich.
(3) Complete neutrino flavor depolarization. In this case, there is
50\% flavor conversion, so $\langle E_{\nu_e}\rangle\to(\langle
E_{\nu_e}\rangle+\langle E_{\nu_\tau}\rangle)/2\approx18$ MeV, which
implies that $Y_e\approx1/(1+\langle E_{\bar\nu_e}\rangle/\langle
E_{\nu_e}^\prime\rangle)\approx0.53,$ too proton rich for $r$-process
nucleosynthesis.
In fact, we must have $Y_e<0.5$ to get {\it any} $r$-process
nucleosynthesis. This is a conservative limit, since a {\it good}
$r$-process requires $Y_e\leq0.45$. Now consider a fourth case,
(4) Partial flavor depolarization. For 35\% conversion of $\nu_\tau$
into $\nu_e$, we have $Y_e\approx 0.5$. For 30\% conversion of
$\nu_\tau$ into $\nu_e$, we have $Y_e\approx 0.49,$ realistically too
large to give acceptable $r$-process nucleosynthesis. In fact, to get
$Y_e<0.45$ we need to demand that there had been less than 15\% flavor
conversion.
Consider some condition along the $Y_e=0.5$ line in Fig. 2 of
Ref. [5]. For example, consider $E_\nu=25$ MeV and $\delta m^2\approx
900$ eV$^2$, for which the density scale height at resonance is
$H\approx0.5$ km. This value of $\delta m^2$ corresponds to
$\sin^22\theta\approx4\times10^{-6}$ on the $Y_e=0.5$ line. In this
case, the conversion probability is
$P({\nu_\tau\to\nu_e})\approx1-\exp\{-0.04\left({\delta m^2/ {\rm
eV}^2}\right)\left({{\rm MeV}/ E_\nu}\right)\left({H/{\rm
cm}}\right)\sin^22\theta\}\approx 25\%$. The neutrino energies around
25 MeV are the most important in terms of leverage on $Y_e$. Note that
fluctuation-induced depolarization at a level of 50\% could
conceivably produce {\it more} conversion than MSW transformation at
$E_\nu=25$ MeV over a considerable part of the region to the {\it
right} of the $Y_e=0.5$ line in Fig. 2 of Ref. [5]. As outlined above,
greater than 30\% flavor conversion will always drive the material too
proton rich for $r$-process.
One can conclude that the random fluctuations will have a quite minor
effect on the neutrino mixing parameters constrained by $r$-process
nucleosynthesis because the absolute increase of the survival
probability rapidly diminishes with increasing survival
probability. To illustrate this, we plot the survival probability as a
function of energy for $\nu_\tau$ with fluctuations of 1\% and 0.5\%
of the local matter density in Figs. 2(a) and 2(b). We choose the
parameters of Fig. 1(b) so that the evolution is nonadiabatic, and the
neutrino energy is chosen to maximize the differential capture rate
[see Eq. (23)]. In Fig. 2, the jagged lines are the solution using
Eq. (9), the solid line is the approximate solution, and the dashed
line is the survival probability in the absence of fluctuations. One
observes that the approximate solution does fairly well in reproducing
the simulation. There is a small increase of about $0.08$ in the
survival probability in Fig. 2(a) where the fluctuations are 1\% of
the local density, but only a very small increase is obtained for
0.5\% noise in Fig. 2(b). If one calculates $Y_e$ for these parameters
in the absence of fluctuations, one obtains $Y_e = 0.5$ indicating
that this point lies on the boundary of the excluded region of the MSW
parameter space [5]. The inclusion of noise at the 1\% level would
decrease $Y_e$ by about 1\%. Random fluctuations of 0.5\% as in
Fig. 2(b) would give a decrease of about 0.4\%. For noise at the
0.1\% level, there will essentially be no change in the excluded
region.
Note that if the neutrino flavor transition is adiabatic, the
fluctuations can increase the survival probability from approximately
zero to one half. The shock reheating epoch occurs at approximately
TPB $\sim0.15$ s, and the relevant scale heights are larger than those
at TPB $>3$ s, which implies a larger resonance width for comparable
values of $\delta m^2$ and $\sin^22\theta$. Hence, $\gamma$ may remain
$\sim 1$ while $\langle(N_e^r)^2\rangle$ is reduced allowing one to
obtain a sizable effect for adiabatic transitions as shown in
Fig. 1(a) for TPB $\approx3$ s with a smaller rms value of the noise.
In Fig. 3 we show the survival probability as a function of neutrino
energy for the values $\delta m^2 = 10^3$ eV$^2$, $\sin^2 2\theta =
10^{-6}$, $\tau_c = 2.762 \times 10^3$ cm. The rms fluctuations are
0.05\% and 0.1\% of the local density in Figs. 3(a) and 3(b),
respectively. In both cases there is a large region where the survival
probability is increased from the value of essentially zero in the
absence of fluctuations. The approximate solution seems to be better
in Fig. 3(b) where the rms value of the fluctuations is larger, but is
reasonably close to the numerical simulation in both cases. For the
case of an adiabatic transition between the more energetic $\nu_\tau$
and less energetic $\nu_e$, the heating rate can be increased by
(30--60)\% [3]. Therefore, the reduction of the transition probability
by random fluctuation effects for an a priori reasonable rms value of
the fluctuations [(0.05--0.1)\%] could be important.
We can estimate the decrease in the heating rate due to the
fluctuations over that from the case of an adiabatic MSW transition as
follows. In the absence of neutrino flavor conversion, the heating
rate per nucleon is approximately given by [13]
\begin{equation}
\dot {\epsilon}_{\nu N} \approx {{L_{\nu}}\over {4 \pi r^2}}
{{\int_0^{\infty} E^3_{\nu} \sigma_{\nu N}
dE_{\nu}/(\exp({E_{\nu}/T_{\nu}}) + 1)} \over
{\int_0^{\infty}E^3_{\nu} dE_{\nu}/(\exp({E_{\nu}/T_{\nu}}) + 1)}} =
\int_0^{\infty} E_{\nu} \sigma_{\nu N} {{d\phi(E_{\nu},T_{\nu})}\over
dE_{\nu}} dE_{\nu},
\end{equation}
where
\begin{equation}
\sigma_{\nu N} \approx 9.6 \times 10^{-44} \left( {E_e \over {\rm MeV}}
\right)^2 {\rm cm^2},
\end{equation}
is the absorption cross section, $L_{\nu}$ is the neutrino luminosity
(assumed equal for all species), $T_{\nu}$ is the neutrino
temperature, $d\phi$ is the differential neutrino flux with respect to
neutrino energy, and $E_e$ is the energy of the produced electron or
positron in Eqs. (17) and (18).
The total heating rate combining contributions from both $\nu_e$ and
$\bar\nu_e$ is
\begin{equation}
\dot {\epsilon}_{\rm tot} = Y_n \dot {\epsilon}_{\nu_e n} + Y_p \dot
{\epsilon}_{\bar{\nu}_e p},
\end{equation}
where $Y_n$ and $Y_p$ are the number fractions of free neutrons and
protons, respectively, with $Y_n + Y_p \approx 1$. Of course, $Y_n$
and $Y_p$ will be set by the same weak interactions responsible for
heating the shock. In the region where neutrino heating dominates, we
can assume
\begin{equation}
{Y_n \over Y_p} \approx {{\lambda_{\bar{\nu}_e p}} \over
{\lambda_{\nu_e n}}},
\end{equation}
where $\lambda_{\nu N}$ is the neutrino capture rate on free nucleons and
is given by
\begin{equation}
\lambda_{\nu N} \approx
\int_0^{\infty} \sigma_{\nu N} {{d\phi(E_{\nu},T_{\nu})}\over dE_{\nu}}
dE_{\nu}.
\end{equation}
An MSW transition between $\nu_e$ and $\nu_\tau$ will increase the
heating rate by a factor of
\begin{equation}
{{\dot {\epsilon}_{tot}^{\prime}}\over {\dot {\epsilon}_{tot}}} = {{
Y_n^{\prime} \dot {\epsilon}_{\nu_e n}^{\prime} + Y_p^{\prime} \dot
{\epsilon}_{\bar{\nu}_e p}} \over {Y_n \dot {\epsilon}_{\nu_e n} +
Y_p \dot {\epsilon}_{\bar{\nu}_e p}}},
\end{equation}
where $\dot {\epsilon}_{\nu_e n}^{\prime}$ is given by
\begin{eqnarray}
\dot {\epsilon}_{\nu_e n}^{\prime} & = &
\int_0^{\infty} P(E_{\nu},\tau_c,\beta ) E_{\nu} \sigma_{\nu N}
{{d\phi(E_{\nu},T_{\nu_e})}\over dE_{\nu}} dE_{\nu} \nonumber \\
& + & \int_0^{\infty}[1 - P(E_{\nu},\tau_c,\beta )] E_{\nu} \sigma_{\nu N}
{{d\phi(E_{\nu},T_{\nu_{\tau}})}\over dE_{\nu}} dE_{\nu}.
\end{eqnarray}
We take $Y_n^{\prime}$ and $Y_p^{\prime}$ to be given by Eq. (22) with
\begin{eqnarray}
\lambda_{\nu_e n}^{\prime} & = &
\int_0^{\infty} P(E_{\nu},\tau_c,\beta ) \sigma_{\nu N}
{{d\phi(E_{\nu},T_{\nu_e})}\over {dE_{\nu}}} dE_{\nu} \nonumber \\
& + &
\int_0^{\infty}[1 - P(E_{\nu},\tau_c,\beta )] \sigma_{\nu N}
{{d\phi(E_{\nu},T_{\nu_{\tau}})}\over {dE_{\nu}}} dE_{\nu}.
\end{eqnarray}
In the above equations, $P(E_{\nu},\tau_c,\beta)$ is the survival
probability of $\nu_\tau$ (or $\nu_e$).
We have calculated Eq. (24) for the cases of random fluctuations with
an rms value of $0.05 \%$ and $0.1 \%$ of the local matter density,
and for the case without fluctuations. We present the results in Table
I. For {\it each} set of values for $(\delta m^2,\ \sin^2 2\theta)$,
we have taken $\tau_c = ({{\delta m^2 \sin 2\theta} / {4 E_{\rm
peak}}} )^{-1}$, where $E_{\rm peak} = 5 T_{\nu_{\tau}}$, the energy
for which the integrand is approximately maximized. We have taken the
temperatures to be $T_{\nu_e} = T_{\bar{\nu}_e} \approx 5$ MeV and
$T_{\nu_{\tau}} \approx 7$ MeV. The values in Table I were calculated
using the approximation in Eq. (13) rather than the ``exact''
numerical simulation, since we are interested in estimating the size
of the difference in neutrino flavor conversion efficiency for the
cases of smooth and noisy density distributions, not on a particular
model of the fluctuations which may or may not obey Eqs. (6) and (7).
{}From Table I, one sees that the noise can reduce the heating rate by
up to 45\% from the adiabatic MSW case.
\section{DISCUSSION AND CONCLUSIONS}
To our knowledge, there is no consensus on the size of possible matter
density fluctuations in post-core-bounce supernovae. Although we have
taken the rms size of the fluctuations to be a constant fraction of
the local matter density, it may very well increase or decrease with
decreasing density, or the size could depend on the distance from the
shock. Similarly, the correlation lengths we have used were chosen to
give the maximal reduction in the MSW transition probability and were
independent of the density and the distance from the shock. If, for
instance, the correlation length increased with decreasing density,
the effect of the noise in reducing the MSW transition probability
could be enhanced since for fixed $\delta m^2$ and $\sin^2 2\theta$
the correlation length of maximal effect ($\tau_c \sim L_{\rm res}$)
varies linearly with the neutrino energy. Our intention in this paper
is to establish the maximal effect density fluctuations could have if
they can be well approximated by random noise added to the average
density.
For an rms fluctuation value of 1\%, the addition of the noise will
have a slight effect on the $r$-process nucleosynthesis as compared
with the MSW effect in the absence of noise. This can be traced to
the fact that demanding neutron-rich conditions at the site of the
$r$-process eliminates all but the nonadiabatic region of the neutrino
mixing parameter space. The effect of the noise is large in the
adiabatic region only and becomes increasingly negligible as the
neutrino flavor transition becomes nonadiabatic. A 1\% fluctuation
will decrease $Y_e$ from $0.5$ to $\sim 0.495$ for points in the MSW
parameter space on the boundary of the excluded region [5]. However,
we consider this rms fluctuation value to be actually too large for
the relatively quiesent TPB $>3$ s epoch. For an order of magnitude
smaller fluctuation which may be more reasonable, the effect of the
noise is totally ignorable. In other words, noise at the 0.1\% level
will not alter the region of the neutrino mixing parameter space
excluded in Ref. [5] by the MSW effect.
Noise with an rms amplitude of 0.05\% of the averaged local matter
density can lead to significant neutrino flavor decoherence during the
shock reheating epoch at TPB $\sim0.15$ s. In turn, this would lead to
a dimunition of the adiabatic MSW-induced increase in the supernova
explosion energy from $\sim$ (30--60)\% to $\sim 20\%$. The physics of
supernova explosion process is quite complicated and depends on many
factors. In our opinion, if the heating of the stalled shock by an
MSW transition is determined to be a necessary component for a
successful explosion, then a more detailed analysis of the effects of
random matter density fluctuations on the MSW transition would be
warranted.
\vskip .2in
We wish to thank Ray Sawyer for bringing this problem to our attention
and providing many valuable insights. We also want to thank Wick
Haxton and Lincoln Wolfenstein for very useful discussions. This work
was supported by the Department of Energy under Grant
No. DE-FG06-90ER40561 at the Institute for Nuclear Theory, by NSF
Grant No. PHY-9503384 at UCSD, by NSF Grant No. PHY-9314131 and by the
Wisconsin Alumni Research Foundation at UW.
\vfill
\eject
|
\section{Introduction}
\label{u1}
Directed paths in random media (DPRM) \cite{KardarRev} are
simple realizations of glassy systems \cite{glass1,glass2}.
Some examples are pinned flux lines (FL) in high-$T_c$ superconductors,
and domain walls in random field and random bond Ising models.
In thermal equilibrium,
a magnetic FL is pinned by defects (oxygen impurities, grain
boundaries, etc.) in the superconductor which lower
its energy \cite{expreview}. The resulting elastic distortions are
limited by the line tension which
opposes the bending of the line. This comptetition leads to a free energy
landscape for the FL which is rather complicated and has many local
minima, i.e. metastable states \cite{BoseGlass}.
When an electric current flows
through the system, the FL feels a Lorentz force perpendicular to its
orientation and to the current direction.
As long as the current is not strong enough to overcome the pinning
forces, the line moves by thermally
activated jumps of line segments between metastable configurations
\cite{KimAnderson,FisherFisherHuse,JoffeVinokur}.
The length of these line segments is estimated by the condition
that the free energy barrier for a jump should be of the same
order as the gain in free energy due to that jump. These dynamics
are believed to be the reason for the nonlinear
voltage-current characteristics found in experiments \cite{expreview}.
Randomly placed impurities in Ising ferromagnets may generate either a random
magnetic field or random exchange couplings \cite{HuseHenley}.
The free energy landscape for domain walls in these systems
is determined by the competition between the pinning energy and
the energy cost (per unit length or area) for creating the wall.
When the system is quenched to a low temperature, the magnetic domains
grow. As for the flux line, the free energy gain due to the motion of a
domain wall segment is expected to be of the same order as the free energy
barrier
which has to be overcome.
Since energy barriers play an important role in the dynamics
of glassy systems, it is essential to know their properties. The scale
of these barriers should grow with observation size $L$ like
a power law $L^\psi$. Usually, it is assumed that the only energy
scale in the system is set by the fluctuations in free
energy which increase as
$L^\theta$, and that therefore $\psi = \theta$
\cite{HuseHenley,JoffeVinokur}. However,
it is also quite possible that the heights of the ridges in
the random energy landscape scale differently from those of the
valleys that they separate, with $\psi>\theta$. Yet another
scenario is that transport occurs mainly along a percolating
channel of exceptionally low energy valleys with $\psi<\theta$.
A first attempt to clarify this situation was taken in Ref.\cite{MDK},
where $\psi = \theta$ was established for a FL moving in 2 dimensions.
Using a combination of analytic arguments and numerical
simulations, lower and upper bounds to the barrier were found.
This argument was then extended to a FL in 3 dimensions \cite{barrier3d},
yielding again $\psi = \theta$.
In this article, we present in more detail the arguments discussed briefly in
these earlier papers, including also systems with
long-range correlated randomness (random-field Ising models) in 2 dimensions.
We obtain in all cases lower and upper bounds to the barrier that scale
as $L^\theta$, except for possible logarithmic factors,
leading to $\psi = \theta$. Furthermore, it is argued that the result
$\psi = \theta$ holds also in higher dimensions, as long as the distribution
of minimal energies decays exponentially. In all cases, the line can
move through the system by encountering energy fluctations of only order
$L^\theta$ around the mean minimal energy. We also show that a line which
initially has a larger energy can reach this region of minimal energies
by crossing barriers of order $L^\theta$ (or smaller).
The outline of the paper is as follows: In section \ref{u2}, we determine
the energy barrier for a FL moving in 2 dimensions. In section \ref{u3}, we
apply the same algorithm to determine the energy barrier to the motion of
domain walls in 2-dimensional random-field Ising systems. In
section \ref{u4}, we study energy barriers for a FL in 3 dimensions and discuss
also the behavior in higher dimensions. In section \ref{u5}, we take a
general look at the energy landscape and show that a line can
move from any initial configuration to a minimal configuration by
going over no barrier higher than $L^\theta$. In section \ref{u6}, we
try to put the definition of energy barriers on a more solid foundation,
and section \ref{u7} argues that the results of the paper can be
generalized to other elastic media with impurities.
\section{Energy barriers for flux lines in 2 dimensions}
\label{u2}
In two dimensions, we represent a DPRM by the following model:
The line is discretized
to lie on the bonds of a square lattice, directed along its
diagonal. Each segment of the line can proceed along
one of two directions, leading to a total of $2^t$ configurations
after $t$
steps. These configurations are labelled by the set of integers $
\left\{ x(\tau)
\right\}$ for $\tau=0,1,\cdots,t$, giving the transverse coordinate
of the line
at each step (clearly constrained such that $x(\tau+1)=x(\tau)\pm
1$). To each
bond on the lattice is assigned a (quenched) random energy equally
distributed
between 0 and 1. The energy of each configuration is the sum of all
random bond energies on the line. Without loss of generality, we set
$x(0)=0$.
Some exact results are known for this model \cite{KardarRev}:
The fluctuations in the free energy at finite
temperature scale as $t^{1/3}$. The meanderings of the transverse coordinate
of the line scale as $t^\zeta$, where $\zeta = 2/3$ is the
roughness exponent. The scaling behavior of the
pinned FL is governed by a zero-temperature fixed point
\cite{HuseHenley} where energy fluctuations scale in the same way.
A FL at low temperatures, and in thermal equilibrium, is likely to
spend most of the time in configurations of minimal energy.
For each endpoint $(t,x)$ with $x = -t, -t + 2, \cdots, t$, there is
a configuration of minimal
energy $E_{min}(x|t)$ which can be obtained numerically in a time of
order $t^2$.
It is known that for $|x| < x_c \propto t^{2/3}$, the function
$E_{min}(x|t)$ behaves as a random walk and is thus asymptotically Gaussian
distributed \cite{KardarRev,HuseHenleyFisher}.
Since beyond the interval $[-x_c,x_c]$ the energy of minimal paths is
systematically larger,
we consider in this paper only the region $[-x_c,x_c]$.
Fig.\ref{MinimalPaths1} shows minimal paths of length
$ t = 256 $ to endpoints between $x = -96$ and $x = +96$.
We want to find the energy barrier that
has to be overcome when the line is moved from an initial minimal
energy
configuration $\{x_i(\tau)\}$ between $(0,0)$ and $(t,-x_f)$ to a final
configuration $\{x_f(\tau)\}$ between
$(0,0)$ and $(t, +x_f)$, with $x_f \equiv x_f(t) \le x_c$.
The only elementary move allowed is flipping
a kink along the line from one side to the other (except at the end point).
Thus the point $(\tau, x)$ can
be shifted to $(\tau, x \pm 2)$. Each route from the initial to the
final configuration is
obtained by a sequence of such elementary moves. For each sequence,
there
is an intermediate configuration of maximum energy, and a barrier
which
is the difference between this maximum and the initial energy. In a
system at
temperature $T$, the probability that the FL chooses a sequence
which crosses a barrier of height $E_B$ is proportional to
$\exp(-E_B/T)$, multiplied by the number of such sequences. We assume
that, as is the case for the equilibrium DPRM, the ``entropic''
factor of the number of paths does not modify scaling behavior. Thus
at sufficiently low temperatures, the FL chooses the optimal sequence
which has to overcome the least energy, and the overall barrier is
the minimum
of barrier energies of all sequences.
Since the number of elementary moves
scales roughly as the area between the initial and final lines, the
number of possible
sequences grows as $t^{xt}$. This exponential growth makes
it practically impossible to find the barrier by examining all
possible sequences,
hampering a systematic examination of barrier energies. Rather than
finding
the true barrier energy, we proceed by placing upper and lower bounds
on it.
The lower bound was given in Ref.\cite{HwaFisher}, and is obtained
as follows:
The endpoint of the path has to visit all sites $(t,x)$ with $|x|
\leq x_f$, and the
energy of any path ending at $(t,x)$ is at least as large as
$E_{min}(x|t)$. Therefore the barrier
energy cannot be smaller than $\max[E_{min}(x|t) - E_{min}(-x_f|t)]$
for ${x \in [-x_f, x_f]}$.
When $x_f$ is sufficiently small, the probability distribution of
this lower bound
is identical to that of the maximal deviation of a random
walk of length $x_f$ \cite{Mikheev}.
The latter is a Gaussian distribution with a mean value proportional to
$\sqrt{x_f}$, and a variance scaling as $x_f$. This growth
saturates for
$x_f$ of the order of $t^{2/3}$, leading to the scaling behaviors,
\begin{eqnarray}
\left\langle {E_-^{(sr)}} (t,x) \right\rangle & = & t^{1/3} f_1^{(sr)}(x /
t^{2/3}),\qquad \text{ and}
\nonumber \\
{\rm var}(E_-^{(sr)})& = & t^{2/3} f_2^{(sr)}(x / t^{2/3}) ,
\end{eqnarray}
for the lower bound and its variance.
The functions $f_1^{(sr)}(y)$ and $f_2^{(sr)}(y)$ are proportional to
$\sqrt{y}$
and $y$ for small $y$, respectively, and go to a
constant for $y = O(1)$. Our simulation results for systems with $t=$
256, 512,
1024, 2048, and 4096 confirm this expectation. Fig.\ref{Barriers1} shows
the scaling functions $f_1^{(sr)}(y)$ and $f_2^{(sr)}(y)$ for different $t$,
and
the collapse is quite
satisfactory. However, the initial growth proportional to $ \sqrt{x_f}$, is
not clearly seen at these sizes.
To obtain an upper bound for the barrier, we specify an explicit
algorithm for
moving the line from its initial to final configuration. This is
achieved by
finding a sequence of intermediate steps. It is certainly
advantageous to keep the
intermediate paths as close to minimal configurations as possible. We
therefore proceed in the following way:
We first find the minimal paths connecting $(0,0)$ to the points
$(t - 1, x)$ with $x_i(t - 1) < x < x_f(t - 1)$, and we add a
last step to the left (from $(t - 1, x)$ to $(t, x - 1)$).
If $x_i(t - 1) > x_i(t) = -x_f$, we then move the point $(t, -x_f)$
to $(t, -x_f + 2)$. Now the path has the same endpoint as the first
intermediate minimal path. We then move the path to this first
intermediate configuration (the precise prescription will be
given below), and then we move again the endpoint. This procedure
is repeated, until the path reaches its final configuration.
At each step, we obtain a local barrier path which separates two
neighboring minimal
configurations. The overall barrier is of course the one with the
highest energy.
While it may occasionally be possible to go to the next intermediate
configuration in a single elementary move (as defined above), this
is
generally not the case. Intermediate minimal paths with the same
endpoint may be
quite far apart at coordinates $\tau < t$. The reason is simple:
suppose the
random potential has a large positive fluctuation, a ``mountain.''
The minimal
energy paths will then circumvent this region by going to its right
or left. The last
path going to the left and the first one going to the right have
almost the same energy.
They form a loop which can be quite large and is likely to enclose the
barrier when
both paths separate already at small $\tau$. Such loops have been
conjectured \cite{JoffeVinokur,FisherFisherHuse} to play an important
role
in the low-temperature dynamics of the DPRM. Since the transverse
fluctuations of a minimal path of length $t$ grow as $t^{2/3}$, we
expect the
lateral size of these loops to also be of this order.
The algorithm for moving a line of length $t = 2^n$ from an intermediate
configuration $\{x_1(\tau)\}$ to another
one $\{x_2(\tau)\}$, with $x_2(t) = x_1(t)$ is as follows:
If $x_2(\tau) \le x_1(\tau) + 2$ for all $\tau$, we can choose a
sequence of
elementary moves such that at most two bonds of the line are not on
one
or the other minimal path, leading to a barrier of order 1 between
the two.
If $x_2(\tau) > x_1(\tau) + 2$ for some $\tau$, the two paths enclose
a loop. We then
consider the points ${(t / 2 - 1, x)}$ with $x_1(t / 2 - 1) < x <
x_2(t / 2 - 1)$. For each of
these points, we find a minimal segment of length $t / 2 - 1$, connecting
the point ${(t / 2 - 1, x)}$ to $(0,0)$ by a minimal path,
and we take a final step to the left from
$(t / 2 - 1, x)$ to $(t / 2, x - 1)$.
In the same way, we connect the
points ${(t / 2 + 1, x)}$ with $x_1(t / 2 + 1) < x <
x_2(t / 2 + 1) $ to $x_1(t))$ via minimal paths
and add a first step to the right from $(t / 2, x - 1)$ to $(t / 2, x)$.
Two such segments form together an almost minimal path of length $t$,
constrained to go through intermediate points at $t / 2$
and $t/2 \pm 1$. We next move the line $\{x_1(\tau)\}$
stepwise through
this sequence of almost minimal paths. If $x_i(t / 2 - 1) = x_i(t / 2) - 1$,
we first move the upper segment. If $x_i(t / 2) = x_i(t / 2 + 1) + 1$,
we then move the lower segment. Then we move the middlepoint. We
continue by repeatedly moving the upper segment, the lower segment, and the
middle point, until the final configuration $\{x_2(\tau)\}$ is reached.
(If the length of the line is different from $2^n$,
we might have to choose the upper segment to be larger
by 1 than the lower segment, or vice versa.)
The prescription for moving the segments of length $t / 2$ is
exactly the same as for
paths of length $t$: If the distance between two consecutive
configurations is larger than
2 for some $\tau \in [0, t / 2]$, we consider the points at ${(t /
4 \pm 1, x)}$ in between the
two, and find minimal paths of length $t / 4 - 1$ connecting them to the
initial and final points, and add a step to the middle points.
Next we attempt to move segments of length $t / 2$ by
repeatedly moving
line portions of length $t / 4$. In some cases, when the energy
barrier is large, it
is necessary to proceed with this construction until the cutoff scale of
$t / 2^{n - 1} = 2$ is
reached. Thus, at each intermediate configuration, the line is
composed of one
segment of length $t/2$, one of length $t/4$, etc; ending
with two smallest pieces
of length $t / 2^m$ (equal to 2 in the worst case). The barrier
path is the intermediate configuration with highest energy.
Fig.\ref{MinimalPaths1} shows the barrier paths resulting from the above
construction.
We now estimate the barrier energy resulting from the above
construction.
Each intermediate path is composed of segments of minimal paths with
constrained
endpoints, and we would like to find the probability distribution for
the highest energy.
Constraining the endpoints of a minimal path of length $\tau$
typically increases
its energy by $E_-^{(sr)}(\tau)\propto \tau^{1/3}$. A subset of these
intermediate paths
(those that cross the largest mountains) have constraints imposed on
segments
of length $t$, $t/2$, $t/4$, and all the way down to unity. The
number of paths in this
subset (henceforth referred to as candidate barriers) grows as
$N_c(t)\propto t^\alpha$, with $1<\alpha<1+2/3$.
The lower limit comes from noting that for each loop of size $2^m$ there
exist at least two loops of size $2^{m-1}$,
one in the upper and one in the
lower half of the parent loop, thus $N_c\geq t$.
The upper limit comes from the
total number of intermediate configurations
that grows as $tx_f$. The energy of each candidate
barrier path is obtained in a manner similar to that of the lower bound:
Instead of finding
the maximum of a random walk of length $x_f\propto t^{2/3}$, we now
have to examine
the sum of the maxima for a sequence of shorter and shorter random
walks added
together. The mean value of this sum is related to the convergent
series,
\begin{eqnarray}\label{meanEc}
&&\left\langle E_c^{(sr)}(t) \right\rangle=\nonumber\\
&& = \left\langle
E_-^{(sr)}(t)+ 2\, E_-^{(sr)}(t/2)+ 2\, E_-^{(sr)}(t/4)+\cdots \right\rangle
+A\ln(t) \nonumber\\
&&=\left\langle
E_-^{(sr)}(t) \right\rangle\left(1 + 2\, (2^{-1/3} + 2^{-2/3} +
\cdots)\right) +A \ln(t) + B
\nonumber \\
&& \simeq \left\langle E_-^{(sr)}(t)\right\rangle
\left(-1 + 2(1-2^{-{1/3}})^{-1} \right)
+ A\ln(t) + B \nonumber\\
&&= 8.69... \left\langle E_-^{(sr)}(t) \right\rangle+A\ln(t) + B.
\end{eqnarray}
The correction term $A\ln(t)$, is explained as follows:
Each segment of length $2^m$ is composed of a minimal path of length
$2^m - 1$ and one step which has a random energy (the final or initial step,
depending on whether the segment lies in the upper or lower half of a loop).
So the energy of the segment is equal to the energy $E_-^{(sr)}(2^m)$
of a minimal path of length $2^m$, plus a constant of order 1. Since a
candidate barrier has $n = \ln(t) / \ln(2)$ segments, these constants
add up to $A \ln(t)$, with $A$ of the order of unity.
The constant $B$ in Eq.(\ref{meanEc}) accounts for the
breakdown of the scaling form of the energy increase for small
loops.
The mean angle of the smallest loops (of size 2)
approaches the $45^\circ$
limit; their mean energy growing as $0.5 t_m$. For the larger loops,
the angle
$t_m^{2/3} / t_m$ is small and the energy is $0.23 t_m$. A finite
value of $m$
acts as a cutoff separating the two limits. The energy difference
per unit length between
small and large paths then leads to the additive constant $B$ (of the
order of unity)
in Eq.(\ref{meanEc}).
The barrier energy is the maximum of the $N_c(t)$ energies of all
candidate
barriers. To find its characteristics, we need the whole probability
distribution for
the energy $E_c^{(sr)}(t)$. Since $E_c^{(sr)}$ is the sum of energies
coming from its minimal
segments, the simplest assumption is to regard the segment
energies as
independent, approximately Gaussian, random variables. We then
conclude that $E_c^{(sr)}(t)$ is also Gaussian distributed with a variance,
\begin{eqnarray}\label{varEc}
&& {\rm var}\left( E_c^{(sr)}(t) \right) \nonumber\\
&&={\rm var}\left( E_-^{(sr)}(t)\right)+
2\,\left({\rm
var}\left( E_-^{(sr)}(t/2)\right)+
\cdots\right) \nonumber\\
&&\simeq 4.40\ldots {\rm var}\left( E_-^{(sr)}(t) \right) \propto t^{2/3}.
\end{eqnarray}
Since the different segments are in fact constructed through a
specific recursive
procedure, their independence cannot be justified. In the worst
case that they are completely dependent, the right-hand side of
Eq.(\ref{varEc}) has to be multiplied by $n = \log_2(t)$. Since our
numerical results show no evidence for such a logarithmic factor, we shall not
consider it any further.
It can be checked easily that (for large $N$), the maximum of $N$
independent Gaussian variables of mean $a$ and variance $\sigma^2$, is a
Gaussian of mean
$a+\sigma\sqrt{2\ln N}$ and variance $\sigma^2/(2\ln N)$ \cite{Galambos}.
Since the $N_c(t)$
candidate barriers have large segments in common, their energies are
not
independent. We can approximately take this into account by
assuming a subset of them as independent, leading to $N\propto
t^{\alpha'}$ for
some $\alpha' < \alpha$. We thus obtain the following estimates for
the mean upper
bound in barrier energy,
\begin{eqnarray}\label{meanE+1}
\left\langle E_+^{(sr)}(x,t) \right\rangle&=& \left\langle E_c^{(sr)}(x,t)
\right\rangle
+\sqrt{2\ln N{\rm var} E_c^{(sr)}(x,t)}\nonumber\\
&\simeq& \left( 1 +\beta^{(sr)} \sqrt{\ln t} \right)t^{1/3} g_1^{(sr)}(x /
t^{2/3}),
\end{eqnarray}
and its variance,
\begin{eqnarray}\label{varE+}
{\rm var}\left( E_+^{(sr)}(x,t) \right)&=&{{\rm var}\left( E_c^{(sr)}(x,t)
\right)\over 2\ln N^{(sr)}} \nonumber\\
&\simeq& {t^{2/3} \over \ln t}g_2^{(sr)}(x / t^{2/3}).
\end{eqnarray}
The functions $g_1^{(sr)}(y)$ and $g_2^{(sr)}(y)$ are proportional to
$\sqrt{y}$
and $y$ respectively, for small $y$,
constant at large $y$, and in general different from those of the
lower bound.
Our numerical simulations indeed confirm the above scaling forms. The
scaling
functions $g_1^{(sr)}(y)$ and $g_2^{(sr)}(y)$ are plotted in
Fig.\ref{Barriers1} for
different values
of $t$, after averaging over 2000 realizations of randomness. The
$\sqrt{\ln(t)}$
factors are essential, as a comparable collapse is not obtained
without them.
In fact, the best fit to $<E_+^{(sr)}(t)>$ is obtained by including the
correction to scaling
term $ \propto <E_-^{(sr)}(t)>$,
and with $\beta^{(sr)} = 1$. The numerics
therefore support the
neglect of correlations, and the assumption of a Gaussian distributed
$E_c^{(sr)}(t)$.
As in the lower bound, the initial scaling proportional to $\sqrt{ x_f}$ is
not clearly seen
for the sizes studied. Since the leading power for the scaling of
the lower and upper
bounds is identical, we conclude that the barrier energies also grow
as $t^{1/3}$.
(It remains to be seen if the logarithmic factors are truly
present, or merely an
artifact of our algorithm.)
\section{Energy barriers for domain walls in 2-dimensional
random field Ising systems}
\label{u3}
In the previous section, we considered random bond
energies which were uncorrelated. The analytic
argument for the upper bound relied on the random-walk behavior
of $E_{min}(x|t)$ in
this situation. Thus, the proof for $\psi
= \theta$ can not directly be extended to other situations,
where the distribution of lower bound energies is not known.
An important example is the case of domain walls in 2-dimensional
random-field Ising magnets. The energy for creating a domain
wall is equal to the cost of flipping all spins
on one side of the interface; in turn proportional to the sum of
all random fields on the flipped spins. There are consequently
long-range correlations in the domain wall energy in the direction
perpendicular to the orientation of the wall\cite{MKJAP}.
We describe the configurations
of the domain wall by essentially the same model as the FL, but assigning
to each bond a random energy with long-range correlations in the
$x$-direction.
These correlations are generated by first selecting for each time $t$, random
numbers $\{r_t^{(-N)}, r_t^{(-N+1)}, \cdots , r_t^{(N-3)},r_t^{(N-1)}\}$
equally
distributed between $-1$ and $1$, where $N$ is (at
least) as large as the largest time occuring in the simulations. To each
bond connecting $(t,x)$ to $(t + 1, x\pm 1)$ we then assign the energy
\begin{displaymath}
{1\over \sqrt{2N}}
\left(\sum_{i = -N}^{x - 1/2 \pm 1/2} r_t^i - \sum_{x + 1/2 \pm 1/2}^{N-1}
r_t^i \right) \, .
\end{displaymath}
Fig.\ref{MinimalPaths2} shows minimal paths of length
$ t = 128 $. Due to the correlations, neighboring bonds have almost
the same energy, and therefore minimal paths tend to have large
parallel portions.
Fig.\ref{walk} shows the minimal energy as function of the endpoint
position for a given realization of randomness, and for $t = 1024$.
This curve is much smoother and has longer correlations
than the corresponding curve in the case of short-range correlated
randomness, where the minimal energy performs a random walk.
The fluctuations in free energy of a line are known to scale as
$t$, and the roughness exponent is $\zeta = 1$\cite{correlated}.
We determined numerically the
distribution function for the minimal energy shown in Fig.\ref{Emin2}.
It is very close to a Gaussian, with no apparent power-law tails.
We will show that, due to this property of the minimal energy distribution,
the lower and upper bounds to the barrier scale in the same way as
the fluctuations in minimal energy.
As in the previous section, we move the line from an initial minimal
energy
configuration $\{x_i(\tau)\}$ between $(0,0)$ and $(t,-x_f)$, to a final
configuration $\{x_f(\tau)\}$ between
$(0,0)$ and $(t, x_f)$.
Since the endpoint of the path has to visit all sites $(t,x)$ with $|x|
\leq x_f$, and since the
energy of any path ending at $(t,x)$ is at least as large as
$E_{min}(x|t)$, the barrier
energy cannot be smaller than $\max[E_{min}(x|t) - E_{min}(-x_f|t)]$.
Since the distribution of minimal energies
decays exponentially and has no power-law tails, we can expect that the
lower bound scales in the same way as the fluctuation of the minimal
energy, leading to
\begin{eqnarray}
\left\langle {E_-^{(lr)}} (t,x) \right\rangle & = & t f_1^{(lr)}(x / t),
\qquad \text{ and}
\nonumber \\
{\rm var}(E_-^{(lr)})& = & t^2 f_2^{(lr)}(x / t) ,
\end{eqnarray}
for the lower bound and its variance.
Our simulation results for systems with $t=$
256, 512,
1024 and 2048 confirm this expectation. Fig.\ref{Barriers2} shows
the scaling functions $f_1^{(lr)}(y)$ and $f_2^{(lr)}(y)$ for different $t$,
and the collapse is quite
satisfactory. However, the initial growth proportional to $ {x_f}$, is
not clearly seen at these sizes.
Fig.\ref{distrlower} shows the distribution of lower bound energies.
It is very close to a (half)-Gaussian of width proportional to $ t$.
An upper bound can be obtained by exactly the same algorithm as before.
The analytic argument made in the previous section, however,
cannot be directly repeated, since the function $E_{min}(x|t)$ is no longer
a random walk in $x$, and since we do not have analytic results for the lower
bound. We can, however, combine analytic arguments with the numerical
results for the lower bound to predict the scaling behavior of the upper
bound. Since the line is always composed of minimal segments, the energy
of a candidate barrier which has segments of all lengths down to the cutoff
is given by
\begin{eqnarray}\label{meanEc2}
&&\left\langle E_c^{(lr)}(t) \right\rangle \nonumber\\
&&=\left\langle
E_-^{(lr)}(t)+2\,\left(E_-^{(lr)}(t/2)+E_-^{(lr)}(t/4)+\cdots\right)
\right\rangle \nonumber\\
&&\simeq 3 \left\langle E_-^{(lr)}(t) \right\rangle + A' \ln(t) + B'.
\end{eqnarray}
The origin of the terms $A' \ln(t) + B'$ has been
explained in the previous section
(see paragraph after Eq.(\ref{meanEc})).
Since the energy distribution of the lower bound is
approximately (half-)Gaussian, the energy distribution of
the candidate barriers decays also like a Gaussian.
The upper bound to the barrier energy is the maximum of the energies
of all candidate barriers. In our simulations, we find no evidence
for logarithmic factors, indicating that the number of candidate
barriers increases either very slowly, or not at all, with $t$.
{}From
Fig.\ref{MinimalPaths2} we can see that there is essentially one large
loop over a distance of the order of the length of the path, leading
to only few independent candidate barriers.
As in the
previous section, we find that the maximum of these candidate barriers,
which is the upper bound to the barrier energy,
scales in the same way as the lower bound, i.e.
\begin{eqnarray}\label{meanE+2}
\left\langle E_+^{(lr)}(x,t) \right\rangle&=& \left\langle E_c^{(lr)}(x,t)
\right\rangle
+\sqrt{2\ln N^{(lr)}{\rm var} E_c^{(lr)}(x,t)}\nonumber\\
&\simeq& t\, g_1^{(lr)}(x / t),
\end{eqnarray}
and its variance scales as
\begin{eqnarray}\label{varE+2}
{\rm var}\left( E_+^{(lr)}(x,t) \right)&=&{{\rm var}\left( E_c^{(lr)}(x,t)
\right)/ 2\ln N^{(lr)}} \nonumber\\
&\simeq& {t^{2}} g_2^{(lr)}(x / t).
\end{eqnarray}
Fig.\ref{Barriers2} shows the scaling functions $f_2^{(lr)}$
and $g_2^{(lr)}$.
To summarize the results so far, we have established the relation
$\psi = \theta$
for lines in 2-dimensional systems with short- and long-range correlated
randomness. Since the considerations for
both systems rely strongly on the dimensionality, it is of importance
to look also at a 3-dimensional system, which is physically more relevant.
\section{Energy barriers for a flux line in 3 dimensions}
\label{u4}
In a two-dimensional system, the endpoint of the FL has to move through
all points $(x,t)$ with $x_i < x < x_f$. This property was essential for
the derivation of the lower bound in the previous sections.
A FL which moves in three
dimensions can avoid regions in space which are energetically unfavorable
for some of its segments, and one may therefore
speculate that $\psi<\theta$.
In this section, we first determine numerically a lower bound for the
barrier
energy which scales in the same way as the energy fluctuations,
thus ruling out $\psi<\theta$. Further numerical results
predict that an upper bound scales in the same way,
thus leading to $\psi = \theta$.
The line now lies on the bonds of a cubic lattice, starting
at the origin and directed
along its (1,1,1) diagonal. Each segment of the line can proceed
in the positive direction along
one of the three axes, leading to a total of $3^t$
configurations after $t$ steps, with endpoints lying in the plane
which is
spanned by the points $(t,0,0)$, $(0,t,0)$, and $(0,0,t)$.
A given configuration of the FL is labelled
by vectors $\left\{ \vec x(\tau)
\right\}$ for $\tau=0,1,\cdots,t$, giving the transverse coordinates
of the FL at each step. The
points $\left\{ \vec x(\tau)\right\}$ lie on the
vertices of a triangular lattice. For a given value of $\tau$, they
lie on one of three alternating sublattices.
The minimal energy $E_{min}(\vec x |t)$ can be obtained
numerically in a time of order $t^3$.
The fluctuations in minimal
energy are known to scale as $t^\theta$ with $\theta \simeq 0.24$,
and the roughness exponent for minimal paths
is $\zeta \simeq 0.62$ \cite{AmarFamily,KimBrayMoore}.
The endpoints of the minimal paths with the lowest energy lie
within a distance proportional to $ t^\zeta$ from the origin. Figure
\ref{profile}
shows the minimal energies of paths of length $t = 288$ to
endpoints $\vec x$ with $|\vec x| < O(t^\zeta)$. The highest energy in
this region is represented in white, the smallest energy in black.
The minimal energies are correlated over
a distance of the order of $t^\zeta$. The distribution of minimal
energies is close to a Gaussian and is shown in Fig.\ref{Emin3}.
Similar to a 2-dimensional system \cite{HalpinHealy}
(see also Fig.\ref{Emin2}),
this distribution seems to have a third cumulant since it is not
completely symmetric.
A lower bound to the barrier energy is obtained as follows:
While the line moves from its initial to final
configuration, the transverse coordinates of its endpoint move
between nearest-neighbor
positions on one of the above mentioned triangular sublattices.
When the endpoint is at a position $\vec x$,
the energy of the line is at least as large as the minimal energy
$E_{min}(\vec x |t)$. The maximum of
all these minimal energies along the trajectory of the endpoint,
minus the energy of the initial configuration, certainly bounds
the barrier energy from below. Since we do not know the actual
trajectory of the endpoint, we have to look for the trajectory
with the smallest maximal energy. Only in this way
can we be sure that we have indeed found a lower bound. This
situation is fundamentally different from a 2-dimensional system,
where there is only one possible trajectory for the endpoint.
Provided that the minimal energies $E_{min}(\vec x |t)$ are known,
this lower bound is determined
in polynomial time by using a transfer-matrix method:
We start by assigning to the initial point $\vec x_i$ a ``barrier
energy'' $B(\vec x_i) = 0$, and to all other sites $\vec x$ on the same
sublattice a barrier energy $B(\vec x) = t$,
which is certainly larger than the lower bound resulting
from the algorithm after many iterations. At each step, the
energy $B$ of all sites $\vec x$, except for the initial site, is updated
according to the following rule: Look for the minimum of the energies
$B(\vec x \pm \vec e_i)$ of the 6 neighbors. If this is smaller than
$B(\vec x)$, replace $B(\vec x)$
by this minimum, or by $E_{min}(\vec x|t) - E_{min}(\vec x_i|t)$,
whichever is larger. After a sufficiently large number of
iterations,
which is of the order of the size of the area of interest
(scaling as $t^{2\zeta}$), all possible trajectories to endpoints
within this area have been probed, and the barrier energies $B(\vec x)$ do not
change any more. The energy $B(\vec x_f)$ is then identified
as the lower bound. Figure \ref{lower} shows the lower bound to the
energy barrier for a line with the endpoint moving from the origin
to sites within a distance of the order of
$t^\zeta$, for different values of $t$ and averaged over 500
realizations of randomness. The distance $|\vec x_f - \vec x_i|$ has been
scaled by $t^{\zeta}$, and the energy by $t^{\theta}$. After
this rescaling, all the curves collapse, leading to
the following scaling behavior for the lower bound,
\begin{equation}
\left\langle {E_-} (t,|\vec x_f - \vec x_i|) \right\rangle =
t^{\theta} f_-(|\vec x_f - \vec x_i| / t^{\zeta}).\label{eq1}
\end{equation}
The function $f(y)$ should be proportional to $y^{\theta/\zeta}$
for small $y$. Again, for the simulated system sizes,
this asymptotic scaling is not clearly seen.
For $y > 1$, the scaling form in Eq.(\ref{eq1}) breaks down
since the minimal energy is then a function of the angle $(|\vec x| / t)$.
We conclude that the lower bound to the barrier scales in the same way
as the
fluctuations in minimal energy, and consequently the energy
barrier increases at least as $t^\theta$, leading to $\psi \ge
\theta$.
The distribution $P(E_-)$ of the lower bound energy
for a fixed distance $|\vec x| \propto t^\zeta$ is
shown in Fig.\ref{distribution}.
It appears to be half-Gaussian with width proportional to $t^\zeta$.
The result $\psi \ge \theta$ is not surprising if we note
that an even simpler lower bound is given by $\max(E_{min}(\vec x_f|t)
- E_{min}(\vec x_i|t), 0)$, which evidently
scales as $t^\theta$ since the
distribution function of minimal energies decays exponentially
fast, i.e. has no power-law tails (see Fig.\ref{Emin3}).
To make sure that the scaling of the lower bound found above is not
dominated by the neighborhood of final configurations with particularly
high energies, we repeated the above simulations
by allowing only endpoints with minimal
energies smaller than the initial energy.
This corresponds to a situation where the endpoint of the line only
moves to positions which are energetically more favorable.
The results are
shown in Fig.\ref{lower} and again collapsed by the scaling form
\begin{equation}
\left\langle {\tilde E_-} (t,|\vec x_f - \vec x_i|) \right\rangle =
t^{\theta} \tilde f_-(|\vec x_f - \vec x_i| / t^{\zeta}). \label{eq2}
\end{equation}
As in the previous case, the asymptotic scaling $\tilde f_-(y) \propto
y^{\theta/\zeta}$ for small $y$ cannot be clearly seen. The energy
distribution of the lower bound is again a half-Gaussian of width
proportional to $ t^{\zeta}$ and looks similar
to Fig.\ref{distribution}.
The same scaling behavior is also found when instead of the
optimal trajectory for the endpoint, the shortest trajectory
(a straight line) is chosen. In this case, the mean of the barrier
energy $E_0$ has the scaling form
\begin{equation}
\left\langle {\tilde E_0} (t,|\vec x_f - \vec x_i|) \right\rangle
= t^{\theta} f_0( |\vec x_f - \vec x_i|/ t^{\zeta})\label{eq3}
\end{equation}
(see Fig.\ref{lower}), again with a half-Gaussian
distribution of width proportional to $t^\zeta$.
This, of course, does not represent a lower bound to the true
barrier, but it will be important for the determination of
an upper bound below, and is therefore included here.
The result $\tilde E_- \propto t^\theta$ (Eq.(\ref{eq2}))
can be explained from the exponential
tails of the distribution of minimal energies:
If we asume that the
endpoint of the line moves only in valleys of particularly low
energy, we can successively remove all sites with the largest
minimal energy from the set of possible endpoints, until the connectivity
over the distance $t^\zeta$ breaks down.
The remaining endpoints form
percolation clusters, and their density is given by the
corresponding percolation threshold. (This is analogous to random resistor
networks describing the hopping resistivity for strongly localized electrons.
The resistance of the whole sample is governed by the critical resistor
that makes the network percolate \cite{randomresistor}.) Since the occupied
sites are
correlated over the distances considered, the value for the
threshold is different from the site percolation limit of 0.5 in
an infinite triangular lattice with no correlation between occupied
sites. But for the present purpose, it is sufficient to know that
this threshold is finite, and that therefore a finite percentage
of all sites are below threshold. Since the distribution of
minimal energies decays rapidly, its tail cannot contain a finite
percentage of all sites. We conclude that the threshold is
within a distance of $t^\theta$ from the peak, and therefore that
the energy fluctuation on the percolation cluster, and
consequently the lower bound for the barrier, are proportional to
$t^\theta$.
We now proceed to construct an upper bound to the energy
barrier. To this purpose, we specify a sequence of
elementary moves which take the line from its initial to
final configuration.
The only elementary move allowed is flipping a
kink along the line. Thus the point $(\tau, \vec x)$ can
be shifted to $(\tau, \vec x \pm \vec e_i)$, where $\pm \vec e_i$ are
the six vectors which connect a vertex in the triangular lattice
to its nearest neighbors within the same sublattice. The algorithm is
similar to the one in 2 dimensions:
First, we choose
a sequence of endpoints connecting the initial to the
final endpoint which is as short as possible. Then, we draw
all the minimal paths leading to these endpoints, and attempt to
move the line through them sequentially. If two consecutive
minimal paths have nowhere a distance
larger than 1 (measured in units of $|\vec e_i|$), we can choose a
sequence of elementary moves such that at most two bonds of the
line are not on one or the other minimal path, leading to a
barrier of order 1 between the two. If the distance is larger than
1, we proceed essentially in the same way as in 2 dimensions, i.e. we
consider the
midway points $(t /2, \vec x_i)$ which connect both
lines via the shortest possible
trajectory $x_i(t/2)$ (if there are several possibilities, we choose one
at random). For each of
these points, we find two minimal segments of length $t/2$
connecting on one side
to $(0, \vec x)$ and on the other to either $(t, \vec x_1(t))$ or
$(t, \vec x_2(t))$. Then we move the line by repeatedly moving segments
of length $t/2$, etc.
The energy of a candidate barrier is then given by
\begin{eqnarray}
&&\langle E_c^{(3d)}(t,t^\zeta)\rangle \nonumber\\
& &\simeq \langle E_0(t,t^\zeta\rangle) \,
(1 + 2\, \left((1/2)^\theta +
(1/4)^\theta + \cdots \right)\nonumber\\
& &= \langle E_0(t,t^\zeta)\rangle\, \left( -1 + 2/(1-(1/2)^\theta)\right)
\nonumber\\
& & = 12.0\ldots\, \langle E_0(t,t^\zeta)\rangle . \label{Ec3d}
\end{eqnarray}
In principle, one should add correction terms similar to those in
Eqs.\ref{meanEc} and \ref{meanEc2}. However, these corrections
are subleading with respect to $t^\theta$ and will be neglected.
As mentioned in section \ref{u3}, we cannot rule out that the
energies of minimal segments are independent from each other.
In the worst case, where they are completely dependent, Eq.(\ref{Ec3d})
has to be multiplied by $\sqrt{\log_2 t}$. This may result in an additional
factor proportional to $ \sqrt{\ln t}$ in the upper bound, but does not
otherwise affect any of our conclusions.
The number of independent candidate barriers increases with some
power in $t$. Since their energy distribution decays like a Gaussian, we
can take their maximum in the same way as before, and
we finally obtain the following estimate for
the upper bound,
\begin{eqnarray}
&& E_+^{(3d)}(t,|\vec x_f - \vec x_i|) = \nonumber \\
&& = \langle E_c(|\vec x_f - \vec x_i|,t) \rangle
+\sqrt{2\ln N}{\rm var} E_c(|\vec x_f - \vec x_i|,t)
\nonumber \\
& \simeq & \left(\sqrt{\ln t} \right)t^{\theta} f_+
(|\vec x_f - \vec x_i| / t^{\zeta})\, .
\end{eqnarray}
We have thus shown that the energy barrier encountered
by a FL moving in a $2d$ or $3d$ random medium has an upper and a lower
bound which both increase as $t^\theta$, except for
logarithmic factors. It thus follows that the
barrier itself scales as $t^\theta$, confirming
the hypothesis $\psi = \theta$.
Since the arguments are mainly based on the
exponential tails of the minimal energy distributions, it is expected
that the result $\psi = \theta$ holds also in higher dimensions.
The only requirement is that the tails in the distributions of minimal
energies still decay sufficiently rapidly.
\section{Barriers to far from minimal configurations}
\label{u5}
In the previous sections, we discussed energy barriers
which have to be overcome by a line
moving between minimal energy configurations. We showed that such lines
can stay in an energy interval $\langle E_{min} \rangle \pm \text{const }
t^\theta$. However, a line may initially have an
energy which is much larger.
The initial configuration of a FL penetrating the system may
be straight and parallel to the external magnetic field. If a system
is cooled down from high temperatures, configurations of the FL
are random walks of roughness exponent
$\zeta = 1/2$. An initial configuration with roughness exponent
$\zeta = 1$ is found for FLs driven close to
a depinning transition\cite{Deniz}.
If the temperature is low (as we always
assume in this paper), the line then relaxes
to some metastable state.
The FL will ultimately reach a configuration of
minimal energy, only if it is not separated from it by abnormally
high barriers. We therefore show in this section that the line can
reach the minimal energy region by going
only over barriers which are not larger than order of $t^\theta$.
We specify an algorithm for moving a line of length $t = 2^n$
{\it from any initial configuration}
to one of minimal energy. The algorithm
is similar to that presented in the previous sections, and leads
to barriers of the order of $t^\theta$.
First, we
assume that its initial roughness is not larger than that
of minimal energy paths.
Let $\{x_n(\tau)\}$ for $\tau = 0,\cdots,t$ be the initial configuration of
the line, and $\{x_0(\tau)\}$ a minimal energy configuration with $x_n(0) =
x_0(0)$ and $x_n(t) = x_0(t)$. We then define a sequence of
paths $\{x_m(\tau)\}$, $m = 1,\cdots,n-1$, which are
constrained to go through the points $x_n(k t/2^m)$ for $k=0,1,\cdots,2^m$
and are composed of $2^m$ minimal segments of length $t / 2^m$.
The energy of such a segment is smaller than the energy of any other
piece of a path with larger $m$ which has the same endpoints as the
segment.
We now move the line successively through this sequence of configurations,
going from the largest to the smallest value of $m$. The configurations
$\{x_{m + 1}(\tau)\}$ and $\{x_{m}(\tau)\}$ intersect each other at the points
$\tau = kt/ 2^m$, with $k = 0,\cdots,2^m$. We therefore can move the line from
the configuration $\{x_{m + 1}(\tau)\}$ to $\{x_{m}(\tau)\}$ by
successively moving segments of length $t/2^m$. In many cases,
the segments have to overcome a loop, and then we apply the algorithm
defined previously. In contrast to the previous sections,
these loops do not separate two minimal configurations,
but one minimal segment, and another constrained at its midpoint,
a constellation which occured also in the previous sections
as an intermediate situation. Since we restricted the roughness of the
initial configuration to less than that of minimal paths,
the size of the loops does not exceed $t^\zeta$. The number of independent
candidate barriers within a loop is therefore smaller than, or equal to,
$N_c \propto (2^m)^{\alpha'}$ from previous arguments, where
the exponent $\alpha'$ depends on the model. The energy of each
candidate path is smaller than, or equal to, the energy $E_c(2^m)$, which was
also obtained in the previous sections. The total number of loops is
less than or equal to
$1+2+\cdots+2^{n-1} < 2^n = t$, and the energy of each candidate barrier
is certainly overestimated if we assume that all loops are of size $t$.
We therefore find an upper bound to the barrier which
is the maximum of $t\, t^{\alpha'} = t^{1 + \alpha'}$ candidates
chosen from a distribution $P(E_c(t))$ with
$\langle E_c(t)\rangle \propto t^\theta$, and with a Gaussian tail.
As we saw in section \ref{u3}, such a maximum scales as
$t^\theta \sqrt{\ln t}$.
We therefore have shown that a line can move from any configuration with
roughness exponent less than $\zeta$ to a minimal energy
configuration by crossing barriers which are not larger than
order of $ t^\theta$, provided that the barriers between minimal
configurations scale also as $t^\theta$.
A similar result can be obtained for any initial configuration of the line.
To demonstrate this, let us look at the
configuration $x_t(\tau) = -\tau$, which is as far as possible from a minimal
configuration. We then define a sequence of paths $\{x_m(\tau)\}$, for
$m = t - 2, t - 4,\cdots, 0$, with $x_m(\tau) = -\tau$ for $\tau \le m$,
and connecting the points $(m,-m)$ and $(t,-m)$ by a minimal path. To go
from one configuration to the next one, the line has to overcome a
loop of size no bigger than $ (t-m)^\theta < t^\theta$. There are
consequently proportional to $ t^{1+\alpha'}$
candidates for barrier paths of length between 2 and $t$. We certainly
find an upper bound to the barrier by assuming that all these candidates
have the length $t$, and that their energies
are taken from a (half-) Gaussian distribution of width proportional to
$ t^\theta$.
The upper bound consequently scales as $t^\theta \sqrt{\ln t}$.
\section{Multiple Barriers}
\label{u6}
So far, we tacitly assumed that the activation barrier is given by the
difference of the highest energy encountered by the line and its
initial energy, just as in thermally activated chemical reactions.
This assumption, however, has no solid foundation, since the line
does not simply move over an isolated maximum, but through a random
energy landscape. In addition, it is not at all clear how results
obtained for a point-like particle in a 1-dimensional energy landscape
can be generalized to lines moving in 2- or 3-dimensional systems.
To shed some light at least on the first of these points, we study in
this section a particle in a 1-dimensional energy
landscape at low temperatures. We tilt this landscape by a small
angle to take into account the effect of an external driving force.
Using the Fokker-Planck equation, we calculate the stationary particle
current through this tilted energy landscape. We find that it is not
the difference between the maximal and initial energies,
but the difference between the maximal and
the minimal energies, which determines the activation barrier.
The Fokker-Planck equation for the probability density $P(x,t)$ of an
overdamped particle in one dimension is
\begin{equation}
{\partial P \over \partial t} = \Gamma {\partial\over\partial x}
\left( kT{\partial\over\partial x} + {\partial V(x) \over \partial x}
\right) P(x,t)\, .
\end{equation}
Here, $\Gamma$ is the inverse of the product of
the particle mass and the friction coefficient.
The potential $V(x)$ is the sum of the random potential $V_B(x)$ and
a driving term $-F x $, where $F$ is the constant driving force. Depending
on the boundary conditions, this equation has different stationary
solutions $\partial_t P = 0$. If the boundary is
an infinitely high wall at both ends of the system, we obtain the
equilibrium solution $P(x) \propto \exp(-V(x)/ kT)$, and zero current
$$j = \Gamma \left( kT{\partial\over\partial x} +
{\partial V(x) \over \partial x} \right) P(x,t) = 0\, . $$
We instead look
for a solution where particles enter the system at one end and leave
it at the other end. This solution is most readily found by assuming
periodic boundary conditions, $P(L) = P(0)$ and $V_B(L) = V_B(0)$. This
situation corresponds to a periodic energy landscape which has been tilted,
and where each section of length
$L$ contains the same number of particles. Clearly, this leads to a
stationary flow through
the system, with particles entering a section at one end and leaving it at
the other. In the limit of small driving force $F$, the stationary
current is found by considering only terms up to linear order in $F$ (order
zero gives the equilibrium solution of the untilted system), and is given
by\cite{FokkerPlanck}
\begin{equation}
j = \Gamma F L / \left[\int_0^Le^{-V_B/kT}dx \int_0^Le^{+V_B/kT}dx\right].
\label{maxmin}
\end{equation}
In the limit $T \to 0$, the integrals are dominated by the neighborhoods
of the maximum and the minimum of the potential, leading to
$ j \propto \exp[(V_B^{max} - V_B^{min})/kT]$. This means that the particle
mobility is determined by the difference between the energy maximum and
minimum. This result is plausible since the particles explore all of the
energy landscape and therefore also go down to the valleys and have
to come up all the way again\cite{SLV}.
Generalizing the above arguments to a
line in 2 dimensions is difficult, and we did not
succeed in solving the corresponding Fokker-Planck equation
analytically. Evidently, the line can avoid
configurations with particularly high energy, which seems to justify the
assumption that the barrier is the lowest possible which
separates the initial and the final configurations.
In the light of Eq.(\ref{maxmin}), however, we may need to define the
barrier energy as the difference between the maximum and the minimum,
instead of the difference between the maximum and the initial energy. If so,
we should add to the barrier the difference between the initial energy
and that of the absolute minimum along the trajectory of the line.
We know, however, that the distribution of minimal energies has only
exponental tails, and that therefore both types of barriers
scale in the same way. Consequently, our results do not
depend on the precise definition of the barrier.
To confirm this, we plot in Fig.\ref{upper} the scaling functions
$\bar g_1^{(sr)}$ and $\bar g_1^{(lr)}$ for the
barrier to 2-dimensional lines with either short-range or
long-range correlated randomness, determining the difference between the
maximal and minimal energy of all intermediate configurations of the line.
It can clearly be seen that the barrier energy still satisfies
Eqs.(\ref{meanE+1}) and (\ref{meanE+2}).
\section{Conclusions}
\label{u7}
In this paper we considered various properties of the energy landscape
of one of the simplest realizations of glassy systems. We showed that,
under fairly general conditions, the energy barriers encountered by
a line descending into the region of
minimal energies, or moving within this region, scale in the same way as
the fluctuations in minimal energy. This means, in particular,
that there exist
no metastable configurations which cannot be left by going over
energy barriers smaller than, or equal to,
the fluctuations in minimal energy.
Similar arguments are applicable to interfaces in
random media, like domain walls in 3-dimensional
random bond and random field Ising models. When the interface
moves from an initial to a final configuration, with part of its
boundary fixed, a given boundary point moves along a line. For each
position of this boundary point, there exists a configuration of
minimal energy. The maximum of these minimal energies certainly is
a lower bound to the barrier. If the distribution of minimal energies
has no power-law tails, this lower bound scales in the same way
as the minimal energy fluctuations. An upper bound can be constructed
using a similar algorithm as for the line: Each time the interface (or
a segment of it) has to overcome a loop, we bisect it and repeatedly move
the upper and lower segment through a sequence of minimal configurations.
In this way, the interface is always composed of minimal segments, and
it should scale in the same way as
the lower bound (except for logarithmic factors, and provided that the
lower bound energy distribution has no power-law tails).
Given these results for lines and interfaces, it is likely that they
generally hold
for elastic media with impurities, e.g. for a bunch of flux lines.
The latter situation is certainly of much more physical releance than a
single FL. Our results for a single FL,
may thus have provided a glimpse into the complexity
of the energy landscape of more complicated glassy systems.
Based on results for particles in 1-dimensional energy landscapes,
we also argue that the energy barrier should not be defined with respect
to the initial energy, but to the minimal energy along the trajectory of
the line.
It stills remains a challenge to generalize this argument to lines in 2- or
higher dimensional energy landscapes, and to find a more precise
expression for the response of the line
to a driving force, starting from the Fokker-Planck equation.
\acknowledgements
This work started in collaboration with Lev Mikheev (see Ref.\cite{MDK}).
We have also benefitted from discussions with Alan Middleton, who has
independendly discovered many interesting results pertaining to energy
barriers\cite{Middleton}. BD is supported by the Deutsche
Forschungsgemeinschaft (DFG) under Contract No.~Dr 300/1-1. MK
acknowledges support from NSF grant number DMR-93-03667.
|
\section{Introduction}
A large number of extensions of the SM predict the existence of color
triplet particles carrying simultaneously leptonic and baryonic
number, the so-called leptoquarks. Leptoquarks are present in models
that treat quarks and leptons on the same footing, such as composite
models \cite{comp}, grand unified theories \cite{gut}, technicolor
models \cite{tech}, and superstring-inspired models \cite{rizzo}.
Since leptoquarks are an undeniable signal for physics beyond the SM,
there have been several direct searches for them in accelerators. At
the CERN Large Electron-Positron Collider (LEP), the experiments
established a lower bound $M_{LQ} \gtrsim 45$--$73$ GeV for scalar
leptoquarks \cite{lep}. On the other hand, the search for scalar
leptoquarks decaying into an electron-jet pair in $p\bar{p}$ colliders
constrained their masses to be $M_{LQ} \gtrsim 113$ GeV \cite{ppbar}.
Furthermore, the experiments at the DESY $ep$ collider HERA
\cite{hera} place limits on their masses and couplings, leading to
$M_{LQ} \gtrsim 92-184$ GeV depending on the leptoquark type and
couplings. There have also been many studies of the possibility of
observing leptoquarks in the future $pp$ \cite{fut:pp}, $ep$
\cite{buch,fut:ep}, $e^+e^-$ \cite{fut:ee}, $e\gamma$ \cite{fut:eg},
and $\gamma\gamma$ \cite{fut:gg} colliders.
In this work we study the constraints on scalar leptoquarks that can
be obtained from their contributions to the radiative corrections to
the $Z$ physics. We evaluated the one-loop contribution due to
leptoquarks to all LEP observables and made a global fit in order to
extract the 95\% confidence level limits on the leptoquarks masses and
couplings \cite{nois}. The most stringent limits are for leptoquarks
that couple to the top quark. Therefore, our results turn out to be
complementary to the low energy bounds \cite{leurer,davi} since these
constrain more strongly first and second generation leptoquarks.
The masses and couplings of leptoquarks are constrained by low-energy
experiments, since the leptoquarks induce two-lepton--two-quark
effective interactions, for energies much smaller than their masses
\cite{leurer,davi}. The processes that lead to strong limits are:
$\bullet$ Leptoquarks can give rise to flavor changing neutral current
(FCNC) processes if they couple to more than one family of quarks or
leptons \cite{shanker,fcnc}. In order to avoid strong bounds from
FCNC, we assumed that the leptoquarks couple to a single generation
of quarks and a single one of leptons. However, due to mixing effects
on the quark sector, there is still some amount of FCNC \cite{leurer}
and, therefore, leptoquarks that couple to the first two generations
of quarks must comply with some low-energy bounds \cite{leurer}.
$\bullet$ The analyses of the decays of pseudoscalar mesons, like the
pions, put stringent bounds on leptoquarks unless their coupling is
chiral -- that is, it is either left-handed or right-handed
\cite{shanker}.
$\bullet$ Leptoquarks that couple to the first family of quarks and
leptons are strongly constrained by atomic parity violation
\cite{apv}. In this case, there is no choice of couplings that avoids
the strong limits.
It is interesting to keep in mind that the low-energy data constrain
the masses of the first generation leptoquarks to be bigger than
$0.5$--$1$ TeV when the coupling constants are equal to the
electromagnetic coupling $e$ \cite{leurer}.
The bounds on scalars leptoquarks coming from low-energy and $Z$
physics exclude large regions of the parameter space where the new
collider experiments could search for these particles, however, not
all of it \cite{fut:pp,fut:ep,fut:ee,fut:eg,fut:gg}. Notwithstanding,
we should keep in mind that nothing substitutes the direct
observation.
\section{Effective Interactions and Analytical Expressions}
\label{l:eff}
A natural hypothesis for theories beyond the SM is that they exhibit
the gauge symmetry $SU(2)_L \times U(1)_Y$ above the symmetry breaking
scale $v$. Therefore, we imposed this symmetry on the leptoquark
interactions. In order to avoid strong bounds coming from the proton
lifetime experiments, we required baryon ($B$) and lepton ($L$) number
conservation. The most general effective Lagrangian for leptoquarks
satisfying the above requirements and electric charge and color
conservation is \cite{buch}
\begin{eqnarray}
{\cal L}_{{eff}} & & = {\cal L}_{F=2} ~+~ {\cal L}_{F=0}
\; ,
\nonumber \\
{\cal L}_{F=2} & & = \left ( g_{{1L}}~ \bar{q}^c_L~ i \tau_2~ \ell_L +
g_{{1R}}~ \bar{u}^c_R~ e_R \right )~ S_1
+ \tilde{g}_{{1R}}~ \bar{d}^c_R ~ e_R ~ \tilde{S}_1
+ g_{3L}~ \bar{q}^c_L~ i \tau_2~\vec{\tau}~ \ell_L \cdot \vec{S}_3
~ ,
\label{lag:fer}
\label{eff} \\
{\cal L}_{F=0} & & = h_{{2L}}~ R_2^T~ \bar{u}_R~ i \tau_2 ~ \ell_L
+ h_{{2R}}~ \bar{q}_L ~ e_R ~ R_2
+ \tilde{h}_{{2L}}~ \tilde{R}^T_2~ \bar{d}_R~ i \tau_2~ \ell_L
\; ,
\nonumber
\end{eqnarray}
where $F=3B+L$, $q$ ($\ell$) stands for the left-handed quark (lepton)
doublet, and $u_R$, $d_R$, and $e_R$ are the singlet components of the
fermions. We denote the charge conjugated fermion fields by
$\psi^c=C\bar\psi^T$ and we omitted in (\ref{lag:fer}) the flavor
indices of the couplings to fermions and leptoquarks. The leptoquarks
$S_1$ and $\tilde{S}_1$ are singlets under $SU(2)_L$ while $R_2$ and
$\tilde{R}_2$ are doublets, and $S_3$ is a triplet. Furthermore, we
assumed in this work that the leptoquarks belonging to a given
$SU(2)_L$ multiplet are degenerate in mass, with their mass denoted by
$M$.
Local invariance under $SU(2)_L \times U(1)_Y$ implies that
leptoquarks also couple to the electroweak gauge bosons. To obtain the
couplings to $W^\pm$, $Z$, and $\gamma$, we substituted $\partial_\mu$
by the electroweak covariant derivative ($D_\mu$) in the leptoquark
kinetic Lagrangian:
\begin{equation}
D_\mu \Phi = \left [ \partial_\mu -
i \frac{e}{\sqrt{2} s_W} \left ( W_\mu^+ T^+ + W_\mu^- T^- \right ) -
i e Q_Z Z_\mu
+ i e Q^\gamma A_\mu \right ] \Phi \; ,
\end{equation}
where $\Phi$ stands for the leptoquarks fields, $Q^\gamma$ is the
electric charge matrix of the leptoquarks, $s_W$ is the sine of the
weak mixing angle, and the $T$'s are the generators of $SU(2)_L$ for
the representation of the leptoquarks. The weak neutral charge is $Q_Z
= (T_3 - s_W^2 Q^\gamma)/s_W c_W$.
We employed the on-shell-renormalization scheme, adopting the
conventions of Ref.\ [20]. We used as inputs the fermion
masses, $G_F$, $\alpha_{{em}}$, and the $Z$ mass, and the
electroweak mixing angle being a derived quantity that is defined
through $\sin^2 \theta_W = s_W^2 \equiv 1 - M^2_W / M^2_Z$. We
evaluated the loops integrals using dimensional regularization and we
adopted the Feynman gauge to perform the calculations.
Close to the $Z$ resonance, the physics can be summarized by the
effective neutral current
\begin{equation}
J_\mu = \left ( \sqrt{2} G_\mu M_Z^2 \rho_f
\right )^{1/2} \left [ \left ( I_3^f - 2 Q^f s_W^2 \kappa_f \right )
\gamma_\mu - I_3^f \gamma_\mu \gamma_5 \right ] \; ,
\label{form:nc}
\end{equation}
where $Q^f$ ($I_3^f$) is the fermion electric charge (third component
of weak isospin). The form factors $\rho_f$ and $\kappa_f$ have universal
contributions, {\em i.e.} independent of the fermion species, as well
as non-universal parts:
\begin{eqnarray}
\rho_f & = & 1 + \Delta \rho_{{univ}} +
\Delta \rho_{{non}} \; , \\
\kappa_f & = & 1 + \Delta \kappa_{{univ}} +
\Delta \kappa_{{non}} \; .
\end{eqnarray}
Leptoquarks can affect the physics at the $Z$ pole through their
contributions to both universal and non-universal corrections. The
universal contributions can be expressed in terms of the
unrenormalized vector boson self-energy ($\Sigma$) as
\begin{eqnarray}
\Delta \rho^{LQ}_{{univ}}(s) &=&
-\frac{\Sigma^Z_{LQ}(s)-\Sigma^Z_{LQ}(M_Z^2)}{s-M_Z^2}
+\frac{\Sigma^Z_{LQ}(M_Z^2)}{M_Z^2}
-\frac{\Sigma^W_{LQ}(0)}{M_W^2} - 2 \frac{s_W}{c_W}
\frac{\Sigma^{\gamma Z}_{LQ}(0)}
{M_Z^2} - \chi_e - \chi_\mu
\; ,\\
\Delta \kappa^{LQ}_{{univ}} &=& - \frac{c_W}{s_W}~
\frac{\Sigma^{\gamma Z}_{LQ}(M_Z^2)}{M_Z^2}
- \frac{c_W}{s_W}~
\frac{\Sigma^{\gamma Z}_{LQ}(0)}{M_Z^2}
+\frac{c_W^2}{s_W^2} \left[ \frac{\Sigma_{LQ}^Z(M_Z^2)}{M_Z^2}-
\frac{\Sigma_{LQ}^W(M_W^2)}{M_W^2}\right]
\; ,
\end{eqnarray}
where the factors $\chi_\ell$ are defined below. The leptoquark
contributions to the self-energies can be easily evaluated, yielding
\begin{equation}
{\Sigma}^{V}_{{LQ}}(k^2) = - \frac{\alpha_{{em}}}{4\pi} N_c
\sum_{j} {\cal F}^V_j~
{\cal H} \left ( k^2, M^2\right )
\; , \label{sig:g}
\end{equation}
where $N_c = 3$ is the number of colors and the sum is over all
members of the leptoquark multiplet. The coefficient ${\cal F}^V_j$ is
given by $(Q^\gamma_{j})^2$, $\left ( Q_Z^{j} \right) ^2$, $
-Q^\gamma_{j} Q_Z^{j}$, and $ \left ( T_3^{j} \right )^2/s_W^2$ for $V
= \gamma$, $Z$, $\gamma Z$, and $ W$ respectively. The function
${\cal H}$ is defined according to:
\begin{equation}
{\cal H}(k^2, M^2) = - \frac{k^2}{3} \Delta_M - \frac{2}{9}k^2
- \frac{4 M^2 - k^2}{3} \int^1_0 dx~ \ln \left [
\frac{{ x^2 k^2 - x k^2 + M^2 - i \epsilon}}
{M^2} \right ] \; ,
\end{equation}
with
\begin{equation}
\Delta_M = \frac{2}{4-d} - \gamma_E + \ln(4\pi) - \ln \left (
\frac{M^2}{\mu^2} \right ) \; ,
\label{delta}
\end{equation}
and $d$ being the number of dimensions.
The factors $\chi_\ell$ ($\ell = e$, $\mu$) stem from corrections to
the effective coupling between the $W$ and fermions at low energy.
Leptoquarks modify this coupling, inducing a contribution that we
parametrize as
\begin{equation}
i \frac{e}{\sqrt{2} s_W}~ \chi_\ell~ \gamma_\mu P_L \; ,
\end{equation}
where $P_L$ ($P_R$) is the left-handed (right-handed) projector and
$\ell$ stands for the lepton flavor. Since this correction modifies
the muon decay, it contributes to $\Delta r$, and consequently, to
$\Delta \rho_{{univ}}$. Leptoquarks with right-handed couplings,
as well as the $F=0$ ones, do not contribute to $\chi_\ell$. The
analytical for $\chi_\ell$ due to left-handed leptoquarks in the $F=2$
sector can be found in Ref.\ [14].
Corrections to the vertex $Z f \bar{f}$ give rise to non-universal
contributions to $\rho_f$ and $\kappa_f$. We parametrize the effect of
leptoquarks to these couplings by
\begin{equation}
i \frac{e}{2 s_W c_W} \left [ \gamma_\mu F_{VLQ}^{Zf} - \gamma_\mu \gamma_5
F_{ALQ}^{Zf} + I_3^f \gamma_\mu (1 - \gamma_5) \frac{c_W}{s_W} ~
\frac{\Sigma^{\gamma Z}_{LQ}(0)}{M_Z^2} \right ] \; ,
\end{equation}
where for leptons ($\ell$) and leptoquarks with $F=2$
\begin{equation}
\begin{array}{ll}
F^\ell_{VLQ}= & \pm F^\ell_{ALQ}= \frac{g_{LQ,X}^2}{32 \pi^2} N_c
{\displaystyle \sum_{j, q} }
{M^{j}_{\ell q}}^\dagger M^{j}_{q\ell} \\
& \left\{ \frac{g^q_X}{2}
- s_W c_W Q_Z^{j}- \left (g_X^q + 2 s_W c_W Q_Z^{j} \right )~
\frac{M^2 - m_q^2}{M_Z^2}
\left [ - \frac{1}{2} \ln \left ( \frac{M^2}{m_q^2} \right )
+ \bar{B_0} ( 0, m_q^2,M^2 ) \right ] \right. \\
& + 2 s_W c_W Q_Z^{j} \frac{M^2 - m_q^2 - \frac{1}{2} M_Z^2}{M_Z^2}
\left [ - \ln \left ( \frac{M^2}{m_q^2} \right ) + \bar{B_0}
( M_Z^2, M^2, M^2) \right ] \\
& + g_X^q \frac{M^2-m_q^2 - \frac{1}{2} M_Z^2}{M_Z^2} \bar{B_0}
(M_Z^2, m_q^2, m_q^2 ) + g^{\ell}_X \bar{B_1} (0, m_q^2, M^2) \\
& + \left [ g_{-X}^q m_q^2 + g_X^q \frac{(M^2-m_q^2)^2}{M_Z^2}
\right ] C_0 (0, M_Z^2, 0, M^2, m_q^2, m_q^2 ) \\
& \left. - 2 s_W c_W Q_Z^{j} \frac{(M^2-m_q^2)^2 + m_q^2 M_Z^2}{M_Z^2}
C_0 (0, M_Z^2, 0, m_q^2, M^2, M^2) \right\} \; ,
\end{array}
\label{z:ll}
\end{equation}
where the $+$ $(-)$ corresponds to left- (right-) handed leptoquarks
and $g_{L/R}^f = v^f \mp a^f$ with the neutral current couplings being
$a_f = I_3^f$ and $v_f = I_3^f - 2 Q^f s_W^2$. $M^{j}_{q \ell}$
summarizes the couplings between leptoquarks and fermions. The
functions $B_1$, $C_0$, $C_{00}$, and $C_{12}$ are the
Passarino-Veltman functions \cite{passa}. We used the convention
$X=L,R$ and $-L=R$ ($-R=L$). We also defined
\begin{eqnarray}
B_0 (k^2, M^2, {M^\prime}^2) & \equiv & \frac{1}{2}\Delta_M+
\frac{1}{2} \Delta_{M'} + \bar{B_0}
(k^2, M^2, {M^\prime}^2 )
\; , \\
B_1 (k^2, M^2, {M^\prime}^2) & \equiv & - \frac{1}{2} \Delta_M + \bar{B_1}
(k^2, M^2, {M^\prime}^2)
\; ,
\end{eqnarray}
with $\Delta_M$ given by Eq.\ (\ref{delta}). From this last expression
we can obtain the effect of $F=2$ leptoquarks on the vertex $Z q
\bar{q}$ simply by the change $\ell \Leftrightarrow q$. Moreover, we
can also employ the expression (\ref{z:ll}) to $F=0$ leptoquarks
provided we substitute $g_{LQ,X} \Rightarrow h_{LQ,X}$ and $g^q_{\pm
X} \Rightarrow - g^q_{\mp X}$.
With all this we have
\begin{eqnarray}
\Delta \rho^{LQ}_{{non}} & = & \frac{F_{ALQ}^{Zf}}{a_f}(M_Z^2)
\; , \\
\Delta \kappa^{LQ}_{{non}} & = & - \frac{1}{2 s_W^2 Q^f} \left [
F_{VLQ}^{Zf}(M_Z^2) - \frac{v_f}{a_f}~ F_{ALQ}^{Zf}(M_Z^2)
\right ]
\; .
\end{eqnarray}
One very interesting property of the general leptoquark interactions
that we are analyzing is that all the physical observables are
rendered finite by using the same counter-terms as appear in the SM
calculations \cite{hollik}. For instance, starting from the
unrenormalized self-energies (\ref{sig:g}) and the mass and
wave-function counter-terms we obtain finite expression for the
two-point functions of vector bosons. Moreover, the contributions to
the vertex functions $Z f \bar{f}$ and $W f \bar{f^\prime}$ are
finite.
In order to check the consistency of our calculations, we analyzed the
effect of leptoquarks to the $\gamma f \bar{f}$ vertex at zero
momentum. It turns out that the leptoquark contribution to this vertex
function not only is finite but also vanishes at $k^2=0$ for all
fermion species. Therefore, our expressions for the different
leptoquark contributions satisfy the appropriate QED Ward identities,
and leave the fermion electric charges unchanged. Moreover, we also
verified explicitly that the leptoquarks decouple in the limit of
large $M$.
\section{Results and Discussion}
\label{res}
In our analyses, we assumed that the leptoquarks couple to leptons and
quarks of the same family. In order to gain some insight on which
corrections are the most relevant, let us begin our analyses by
studying just the oblique corrections \cite{obli}, which we
parametrized in terms of the variables $\epsilon_1$, $\epsilon_2$, and
$\epsilon_3$. These variables depend only upon the interaction of
leptoquarks with the gauge bosons and it is easy to see that
leptoquarks contribute only to $\epsilon_2$. Imposing that this
contribution must be within the limits allowed by the LEP data,
we find out that the constraints coming from
oblique corrections are less restrictive than the available
experimental limits \cite{lep,ppbar,hera}.
We then performed a global fit to all LEP data including both
universal and non-universal contributions. In Table \ref{LEPdata} we
show the the combined results of the four LEP experiments \cite{sm}
that were used in our analysis. In order to perform the global fit we
constructed the $\chi^2$ function associated to these data and we
minimized it using the package MINUIT. We expressed the theoretical
predictions to these observables in terms of $\kappa^f$, $\rho^f$, and
$\Delta r$, with the SM contributions being obtained from the program
ZFITTER \cite{zfit}. In our fit we used five parameters, three from
the SM: $m_{{top}}$, $M_H$, and $\alpha_s(M_Z^2)$, and two new
ones: $M$, and the leptoquark coupling denoted by $g_{LQ}$.
Furthermore, we have also studied the dependence upon the SM inputs
$M_Z$, $\alpha_{{em}}$, and $G_F$.
\begin{table}
\caption{LEP data}
\label{LEPdata}
\begin{displaymath}
\begin{array}{|l|l|}
\hline
\hline
\mbox{Quantity} & \mbox{Experimental value} \\ \hline
M_Z \mbox{[GeV]} & 91.1888 \pm 0.0044 \\
\Gamma_Z \mbox{[GeV]} & 2.4974 \pm 0.0038 \\
\sigma_{\rm had}^0 \mbox{[nb]} & 41.49 \pm 0.12\\
R_e = \frac{\Gamma({\rm had})}{\Gamma(e^+ e^-)} & 20.850 \pm 0.067 \\
R_\mu = \frac{\Gamma({\rm had})}{\Gamma(\mu^+ \mu^-)} & 20.824 \pm 0.059 \\
R_\tau = \frac{\Gamma({\rm had})}
{\Gamma(\mu^+ \mu^-)} & 20.749 \pm 0.070 \\
A_{FB}^{0e} & 0.0156 \pm 0.0034 \\
A_{FB}^{0\mu} & 0.041 \pm 0.0021 \\
A_{FB}^{0\tau} & 0.0228 \pm 0.0026 \\
A_{\tau}^0 & 0.143 \pm 0.010 \\
A_e^0 & 0.135 \pm 0.011 \\
R_b = \frac{\Gamma(b \bar{b})}{ \Gamma({\rm had})} & 0.2202 \pm 0.0020\\
R_c = \frac{\Gamma(c\bar{c}) }{\Gamma({\rm had})} & 0.1583 \pm 0.0098\\
A_{FB}^{0b} & 0.0967 \pm 0.0038 \\
A_{FB}^{0c} & 0.0760 \pm 0.0091 \\
\hline
\hline
\end{array}
\end{displaymath}
\end{table}
The first part of our analysis consisted of the study of the constraints
on the leptoquark masses and couplings. In order to determine the
allowed region in the $M_{LQ}$--$ g_{LQ}$ plane, shown in Fig.\
\ref{contours} for the different models, we obtained the minimum
$\chi^2_{{min}}$ of the $\chi^2$ function with respect to the
parameters above for each leptoquark model, and we then required that $\chi^2
\leq \chi^2_{{min}} +\Delta \chi^2(2,90\% \hbox{CL})$, with
$\Delta\chi^2(2,90\% \hbox{CL})=4.61$. In this procedure, the
parameters $m_{{top}}$, $M_H$, and $\alpha_s$, as well as the SM
inputs $M_Z$, $\alpha_{{em}}$, and $G_F$ were varied so as to
minimize $\chi^2$. We must comment here that the dependence on
$\alpha_{{em}}$ and $G_F$ is negligible when they are allowed to
vary in their $90\%$ CL range. On the other hand, the variation of
$M_Z$ in the interval $91.18\leq M_Z\le 91.196$ leads to a change on
the allowed values of leptoquarks parameters of at most 1\%.
The contour plots exhibited in Fig.\ \ref{contours} were obtained for
third generation leptoquarks. From this figure we can see that the
bounds are much more stringent for the leptoquarks that couple to the
top quark, {\em i.e.} for $S_{1L(R)}$, $S_3$, and $R_{2L(R)}$, since
their contributions are enhanced by powers of the top quark mass.
Moreover, the limits are slightly better for left-handed leptoquarks
than for right-handed ones, given a leptoquark type, and the curve is
symmetric around $g_{LQ}=0$ since the leptoquark contributions are
quadratic functions of $g_{LQ}$.
The contributions from $\tilde R_2$ and $\tilde S_1$ are not enhanced
by powers of the top quark mass since these leptoquarks do not couple
directly to up-type quarks. Therefore, their limits are much weaker,
depending on $m_{{top}}$ only through the SM contribution, and
the bounds for these leptoquarks are worse than the present discovery
limits unless they are strongly coupled ($g_{LQ}^2 = 4 \pi$).
Moreover, the limits on first and second generation leptoquarks are
also uninteresting for the same reason. Nevertheless, if we allow
leptoquarks to mix the third generation of quarks with leptons of
another generation the bounds obtained are basically the same as the
ones discussed above\footnote{In the case of first generation leptons,
we must also add a tree level $t$-channel leptoquark exchange to
some observables.}, since the main contribution to the constraints
comes from the $Z$ widths.
We next present our results as 95\% CL lower limits in the leptoquark
mass and study the dependence of these limits upon all other
parameters. For this, we minimized the $\chi^2$ function for fixed
values of $\alpha_s$, $M_H$, and $m_{{top}}$ and then required
$\chi^2 (\alpha_s, M_H, m_{{top}})\le \chi^2_{{min}}
(\alpha_s, M_H, m_{{top}})+ \Delta\chi^2 (1,90\% \hbox{CL})$,
with $ \Delta\chi^2(1,90\% \hbox{CL})=2.71$. Our results are shown in
Table \ref{res:top} where we give the 95\% CL limits obtained for a
third generation leptoquark for several values of the coupling
constants $g_{LQ}$ ($=\sqrt{4\pi}$, $1$, and $e/s_W$). The values
given correspond to $m_{top}=175$ GeV and variation of $M_H=60-1000$
GeV and $\alpha_s(M_Z^2)=0.126\pm 0.005$, which is the range
associated to the best values obtained from a fit in the framework of
the SM \cite{sm}. For a fixed value of $m_{{top}}$ and
leptoquark coupling constant, the dependence on $\alpha_s(M_Z^2)$ and
$M_H$ is such that the limits are more stringent as $\alpha_s(M_Z^2)$
increases and $M_H$ decreases. The SM parameters $M_Z$,
$\alpha_{{em}}$, and $G_F$ have been also varied in their allowed
range. However, this did not affect the results in a noticeable way.
\begin{table}
\caption{ Lower limits (95\% CL) for the mass of third generation
leptoquarks in
GeV for different values of the couplings, assuming $m_{{top}} = 175$
GeV, $\alpha_s(M_Z^2) = 0.126\pm 0.005$, and $M_H = 60-1000$ GeV.}
\label{res:top}
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
\hline
$g_{LQ}$ & $S_1^R$ & $S_1^L$ & $S_3$ & $R_2^R$
& $R_2^L$ & $\tilde{S}_1^R$ & $\tilde{R}_2^L$ \\
\hline
$\protect\sqrt{4\pi}$ & 5800--3200 & 6000--3500 & 8000--3700
& 6000--3300 & 6800--3400 & 300--100 &
550--120\\
1 & 1200--550 & 1200--600 & 1700--700 & 1250--600 & 1400--600
& --- & ---\\
${\displaystyle \frac{e}{s_W}}$ & 550--200 & 600--225 & 900--325 &
600-250 & 700-250 & --- & --- \\
\hline
\hline
\end{tabular}
\end{table}
We would like to stress that the large apparent uncertainty associated
with the value of $\alpha_s$ and $M_H$ can be considered somehow
fictitious as the value of $\chi^2_{{min}}$ grows very fast when we
move from the central value $\alpha_s=0.126$, $M_H=300$ GeV what means
that the quality of the fit for the extreme values of these parameters
is rather bad. For instance, $\alpha_s=0.117$, results in a too high
$\chi^2$, even in the context of the SM ($\chi^2_{{min}}>26/12$).
\section{Acknowledgements}
I would like to thank Alan Sommerer for his hospitality. This work
was supported by Conselho Nacional de Desenvolvimento Cient\'{\i}fico
e Tecnol\'ogico (CNPq) and by Funda\c{c}\~ao de Amparo \`a Pesquisa do
Estado de S\~ao Paulo (FAPESP).
|
\section{Introduction}
The production of heavy quarkonia has
been studied quite extensively in deep-inelastic scattering
(see e.g. ref.\cite {Allaea91}) and in hadron-hadron
collisions (see e.g. ref.{\cite{Kowiea94,DualModel}).
In the latter, at sufficiently high energy, the reaction
proceeds through the annihilation of either a quark or gluon in one
hadron with an antiquark or gluon in the other hadron. In principle one
therefore has a probe of the quark and gluon distributions. For the
quarks this process merely supplements the information obtained in
Drell-Yan and deep-inelastic scattering, but for the gluons this is
one of our few direct probes.
There have recently been suggestions that there may be a somewhat
larger violation of charge symmetry in the minority valence quark
distribution in the nucleon than one might naively have expected. At
intermediate $x$ this effect could be as large as 5\%
\cite{Sather92,RoThLo94}. It is important to try to test these ideas in
Drell-Yan reactions \cite{Londea94}.
Here we examine whether such an effect could induce a significant
forward-backward asymmetry in the production of $J/\Psi$ in proton-neutron
collisions. As there has been no explicit calculation of a
difference in the gluon distribution of a proton and neutron
we also comment on this possibility.
In proton-deuteron collisions the forward-backward asymmetry of
$J/\psi$ production is sensitive not only to charge symmetry
violations but also to nuclear modifications of the deuteron
parton distributions. The latter could be investigated through
experiments feasible at RHIC.
\section{$J/\psi$ production in proton-neutron collisions}
The production of charm-anticharm quark pairs
proceeds in leading order in the strong coupling via gluon-gluon
fusion or quark-antiquark annihilation.
The corresponding cross sections for these subprocesses are \cite{Fritsch77}:
\begin{eqnarray}
\hat \sigma_{g g} (c\bar c,M^2) &=&
\frac{\pi \alpha_s^2}{3 M^2} \left[\left( 1+4
\frac{m_c^2}{M^2} + \frac{m_c^4}{M^4} \right) ln\left(\frac{1+\lambda}
{1-\lambda}\right)- \right.\nonumber\\
&&\left. \hspace{1.5cm} \frac{\lambda}{4}
\left( 7+31\frac{m_c^2}{M^2}\right)\right],\\
\hat \sigma_{q\bar q}(c\bar c,M^2) &=&
\frac{8 \pi \alpha_s^2}{27 M^4} \left( M^2 + 2 m_c^2\right)
\lambda,
\end{eqnarray}
where $\lambda=\sqrt{1-4 {m_c^2}/{M^2}}$. The mass of the charm quark
and the produced $c\bar c$ pair is denoted by $m_c$ and $M$ respectively.
The strong coupling constant $\alpha_s$ is calculated at $M^2$
using a QCD scale $\Lambda = 0.177 \,GeV$.
Multiplying the cross sections of the QCD subprocesses
with the parton distributions of beam and target, $f^b_i$ and $f^t_j$, yields
the $c\bar c$ production cross section
\begin{equation}
\frac{d^2\sigma(c\bar c) }{d x_F d M^2} = \sum_{i,j} f^b_i(x_b) f^t_j(x_t)
\frac{\hat \sigma_{ij}(c\bar c,M^2)}{s\,\sqrt{x_F^2 + \frac{4M^2}{s}}}.
\end{equation}
In the center of mass frame Feynman $x_F$ is defined as the
fraction of beam momentum carried by the produced $c\bar c$ pair
and $s$ stands for the squared center of mass energy.
The light-cone momentum fractions of the active beam and
target parton are
\begin{eqnarray}
x_b = \frac{1}{2} \sqrt{x_F^2 + \frac{4 M^2}{s}} + \frac{1}{2} x_F,\\
x_t = \frac{1}{2} \sqrt{x_F^2 + \frac{4 M^2}{s}} - \frac{1}{2} x_F.
\end{eqnarray}
The $J/\psi$ production cross section is proportional to the
charm production cross section, integrated over the invariant mass
of the $c\bar c$ pair \cite{Fritsch77}.
The integration limits are the thresholds for
$c\bar c$ pair production and $D\bar D$ production:
\begin{equation}
\frac{d\sigma(J/\psi)}{d x_F} = F \int_{4 m_c^2}^{4 m_D^2} dM^2\,
\frac{d^2\sigma(c\bar c)}{d x_F d M^2}.
\end{equation}
The factor $F$ specifies the fraction of events in which $J/\psi$ bound
states are formed.
Despite being a simple model, such a description is well
known to describe many features of heavy quark production,
including the dependence of the $J/\psi$ production cross
section on $x_F$ and the beam energy \cite{Kowiea94,DualModel}.
Since it is of importance for our further discussion, we show
in Fig.~1 the separate contributions to the $J/\psi$ production
cross section from gluon fusion and quark-antiquark annihilation.
We take $m_c=1.5\, GeV$ and use the parton distributions
of ref.\cite{Owens91}.
At small $x_F$ gluon fusion is by
far the dominant mechanism, while for $x_F>0.6$ the $q\bar q$
annihilation takes over.
To investigate charge symmetry violation we will now
focus on $J/\psi$ production in proton-neutron collisions.
The corresponding production cross section involves
the following combination of parton distributions
\begin{eqnarray}
\hat \sigma_{q\bar q}&& \hspace{-0.3cm}
\left[u_p(x_b) \bar u_n(x_t) + \bar u_p(x_b) u_n(x_t)\,+\right.\nonumber\\
&&\;\left.
\hspace{-0.3cm} d_p(x_b) \bar d_n(x_t) + \bar d_p(x_b) d_n(x_t)\right]
\,+\nonumber\\
\hat \sigma_{gg} && \hspace{-0.3cm} g_p(x_b) g_n(x_t)
\end{eqnarray}
Here $q_{p/n}$ are the quark distributions of the proton
and neutron respectively and $g_{p/n}$ the corresponding gluon distributions.
In an isospin rotated world $J/\psi$ production in proton-neutron
collisions becomes $J/\psi$ production in neutron-proton collisions,
i.e. the role of beam and target is interchanged.
If charge symmetry were exact, the difference between the corresponding
cross sections would vanish.
Interchanging the role of beam and target is equivalent to a sign
change in $x_F$.
Hence, the difference of the $J/\psi$ production cross sections at
positive and negative $x_F$
\begin{equation}
\Delta\sigma_{pn}(x_F) =
\left. \frac{d\sigma(J/\psi)}{d x_F}\right|_{x_F} -
\left. \frac{d\sigma(J/\psi)}{d x_F}\right|_{-x_F}
\end{equation}
is driven by charge symmetry violations only.
In detail, $\Delta\sigma_{pn}$ contains the following combination
of parton distribution functions:
\begin{center}
\begin{math}
\begin{array}{rlcllcll}
\hat \sigma_{q\bar q}\,\frac{1}{2} \!
& \Big\{\;(\delta u(x_b) \!&-&\! \delta d(x_b))
& (\bar d(x_t) \!&-&\! \bar u(x_t))
& +\\
&\quad (\delta u(x_b) \!&+&\! \delta d(x_b))
& (\bar d(x_t) \!&+&\! \bar u(x_t))
& - \\
&\quad (u(x_b) \!&-&\! d(x_b))
& (\delta \bar d(x_t) \!&-&\! \delta \bar u (x_t))
& -\\
&\quad (u(x_b) \!&+&\! d(x_b))
& (\delta \bar d (x_t) \!&+&\! \delta \bar u(x_t) ) &+
\end{array}
\end{math}
\end{center}
\vspace*{-0.6cm}
\begin{center}
\begin{math}
\quad [q \longleftrightarrow \bar q]\Big\} \,+
\end{math}
\end{center}
\vspace*{-0.4cm}
\begin{equation} \label{Delta_pn}
\hspace*{-3.cm}\hat\sigma_{gg} \,\phantom{\frac{1}{2}}
\Big\{\delta g(x_b) g(x_t) - g(x_b) \delta g(x_t) \Big\}
\end{equation}
We expressed the neutron distributions through the proton ones,
using the definitions $\delta d \equiv d_p - u_n,\,
\delta u \equiv u_p - d_n$ and
$\delta g \equiv g_p - g_n$, and dropping the index ``p''.
First we will consider contributions to $\Delta\sigma_{pn}$ through
charge symmetry violations in the valence distributions. The corresponding
charge symmetry violating parts of the minority and majority
distributions, $\delta d^v = d_p^v - u_n^v$ and
$\delta u^v = u_p^v - d_n^v$ were extensively discussed in
\cite{Sather92,RoThLo94}.
We use the results of ref.\cite{RoThLo94} which were obtained
within the framework of the MIT bag model.
The magnitude of $\delta d^v$ was found to be
similar to that of $\delta u^v$.
As $d^v$ is generally much larger than $u^v$ the fractional change in
$d^v$ is much greater.
This can be easily understood, since one of the major sources of
charge symmetry violation is the mass difference of the residual
di-quark pair, when one quark of the nucleon is hit in a deep-inelastic
scattering process.
For the minority quark distribution the residual di-quark is
$uu$ in the proton and $dd$ in the neutron. Therefore in the difference
$d_p$ - $u_n$ the up-down mass difference enters twice. On the other hand,
for the majority quark distribution, where the residual system
is $ud$, both for the proton and the neutron, there is no contribution
to charge symmetry breaking.
To calculate $\Delta\sigma_{pn}$ we also need to know the
flavor asymmetry of the quark sea, $\bar d - \bar u$,
which enters in the first term of Eq.(\ref{Delta_pn}).
We take a parameterization from ref.\cite{MeThSi91}:
$x (\bar d(x) - \bar u(x)) = A x^{0.5} (1-x)^7$,
where the normalization $A$ is fixed through
$\int dx (\bar d(x) - \bar u(x)) = 0.15$.
Such a parameterization is in good agreement with recently
discovered violations of the Gottfried sum rule \cite{NMC94}.
We normalize $\Delta\sigma_{pn} (x_F)$ through the $J/\psi$ production
cross section in proton-proton collisions and present in
Fig.~2 results for the ratio
$R_{pn}^{J/\psi} = \frac{\Delta\sigma_{pn}}{d\sigma(J/\psi)_{pp}/dx_F}$
at a center of mass energy $\sqrt{s} = 40\,GeV$.
We find $R_{pn}^{J/\psi} \approx - 0.02$ at large
$x_F\sim 0.6-0.7$. At smaller values of $x_F$ (say below $0.5$) the
cross section difference $\Delta\sigma_{pn}$ vanishes.
This happens for two main reasons.
At small $x_F$ gluon fusion yields the major contribution to
$J/\psi$ production and quark-antiquark annihilation
is of little relevance.
Also $\delta d^v - \delta u^v$, which is much larger than
$\delta d^v + \delta u^v$ \cite{RoThLo94}, enters
$\Delta\sigma_{pn}$
in combination with the small flavor asymmetry, $\bar d - \bar u$.
Charge symmetry breaking in the sea quark distributions has
not been calculated yet. Nevertheless its influence on
$\Delta\sigma_{pn}$ can be estimated.
In the MIT bag model the sea quark distributions are dominated by
contributions characterized through
residual four-quark states, after one quark is hit in a
deep-inelastic scattering process \cite{ScSiTh91}.
By analogy with the valence quark case let us assume
that a major source of charge symmetry breaking is
the mass difference of the residual spectator states.
Scattering on an antiquark which carries the same flavor as
the majority quarks leaves a $uuud$ or $dddu$
residual four-quark state in the case of a proton or neutron
target respectively.
In case of an antiquark of minority flavor a $uudd$ residual
state occurs in both cases.
Since the up-down mass difference enters twice in the
first case but is absent in the second, it seems reasonable to
assume $\delta \bar u \gg \delta \bar d$.
In the following we will neglect $\delta \bar d$.
In Fig.~2 we also show $R_{pn}^{J/\psi}$
for various values of $\delta \bar u$ between $0.01\bar u$ and $0.10\bar u$.
Qualitatively we find minor changes to our former result where
only charge symmetry breaking in the valence distributions was taken
into account.
We may therefore conclude that charge symmetry breaking
in the quark distributions contributes to the difference of
$J/\psi$ production in the forward and backward direction only at
large values of $x_F$ ($x_F\gsim 0.6$). Unfortunately this
kinematic region is not accessible at the moment.
In current measurements, using the neutron beam facility at FNAL, one is
restricted to the region $x_F > -0.1$ \cite{Moss}.
However the news is not all bad. Since charge symmetry
violations in quark distributions do not contribute
to $R_{pn}^{J/\psi}$ at small $x_F$, it is an
ideal place to look for charge
symmetry violations in the gluon distributions.
As we can see from Eq.(\ref{Delta_pn}), a (say) $1\%$
charge symmetry violation in the gluon distribution,
$\delta g \sim 0.01 g$, can lead to $|R^{J/\psi}_{pn}|\lsim 0.02$.
No predictions exist up to now for charge symmetry violations
in gluon distributions.
Since a major part of the glue in
hadrons can be viewed as being radiatively generated from
quarks, charge symmetry violations
in the quark distributions may in principle induce similar effects
in the gluon distributions.
However, since for charge symmetry violation in the radiatively
generated glue the small combination $\delta d^v + \delta u^v$
is relevant, a large signal is not expected. If a large signal were seen
it could only be attributed to charge symmetry
violation in the non-perturbative glue, which would certainly be
a surprise.
\section{$J/\psi$ production in proton-deuteron collisions}
In high energy processes deuterons are often used as a
convenient source of neutrons.
Furthermore, in the near future proton-deutron collisions will be
carried out at RHIC at center of mass energies
between $50$ and $375 \,GeV$ and $x_F> -0.5$ \cite{Moss}.
Therefore, at first sight it seems to be a good idea to
investigate charge symmetry violations via $J/\psi$ production
in proton-deuteron collisions.
However, as we will demonstrate below, nuclear modifications of
deuteron parton distributions, which are interesting in their own right,
are most likely to overtake the effects resulting from charge symmetry
violations.
Let us assume that the parton distributions of the deuteron
are related to those in the proton and neutron via
\begin{eqnarray}
q_D(x) &=& (1+\epsilon_q(x)) \,(q_p(x) + q_n(x)), \quad q = u,d,\\
g_D(x) &=& (1+\epsilon_g(x)) \,(g_p(x) + g_n(x)).
\end{eqnarray}
In proton-deuteron collisions the difference $\Delta\sigma_{pD}(x_F)$
of the $J/\psi$ production cross sections measured in the forward and
backward direction, involves the parton distribution functions:
\begin{center}
\begin{math}
\begin{array}{rlllllcll}
\hat \sigma_{q\bar q}\, \Big\{
& \hspace{-0.3cm}\frac{1}{2} \left( 1 + \epsilon_q(x_b)\right) \big[
& \hspace{-0.3cm} (\delta u(x_b) \!&-&\! \delta d(x_b))
& \hspace{-0.3cm} (\bar d(x_t) \!&-&\! \bar u(x_t))
& +\\
&&\hspace{-0.3cm} (\delta u(x_b) \!&+&\! \delta d(x_b))
& \hspace{-0.3cm} (\bar d(x_t) \!&+&\! \bar u(x_t)) \,\big]
& - \\
& \hspace{-0.3cm} \frac{1}{2} \left( 1 + \epsilon_q(x_t)\right) \big[
& \hspace{-0.3cm} (u(x_b) \!&-&\! d(x_b))
& \hspace{-0.3cm} (\delta \bar d(x_t) \!&-&\! \delta \bar u (x_t))
& +\\
&&\hspace{-0.3cm} (u(x_b) \!&+&\! d(x_b))
& \hspace{-0.3cm} (\delta \bar d (x_t) \!&+&\! \delta \bar u(x_t) )\big]
&+ \\
& \hspace{-0.3cm}\left( \epsilon_q(x_t) - \epsilon_q(x_b) \right)
& \hspace{-0.3cm} (u(x_b) \!&+&\! d(x_b))
& \hspace{-0.3cm} (\bar u(x_t) \!&+&\! \bar d(x_t))
& +
\end{array}
\end{math}
\begin{math}
[q \longleftrightarrow \bar q]\Big\} \; +
\end{math}
\end{center}
\vspace*{-0.3cm}
\begin{equation} \label{Delta_pD}
\begin{array}{c}
\hspace*{-1.cm}\hat \sigma_{gg} \,\Big\{
\left(1 + \epsilon_g(x_b)\right)\,\delta g(x_b) \,g(x_t)\,-
\left(1 + \epsilon_g(x_t)\right)\, g(x_b) \,\delta g (x_t)\,+ \nonumber \\
\hspace{0.cm} 2 \left(\epsilon_g(x_t) - \epsilon_g(x_b)\right)
\,g(x_b) \,g(x_t)\,\Big\}
\end{array}
\end{equation}
Clearly not all contributions to $\Delta\sigma_{pD}$ arise from
charge symmetry violation. There are also terms proportional to the
difference of nuclear effects at $x_b$ and $x_t$.
In the following we will estimate their size and
show that they are most
likely to dominate over contributions from charge symmetry violations.
In Fig.~3 we show the light-cone momentum fractions
$x_b$ and $x_t$ at which the beam and target parton distributions
are probed for different $x_F$ and for different center of mass
energies $\sqrt{s}$.
The invariant mass of the produced quark-antiquark pair is varied
over the range $4 m_c^2 < M^2 < 4 m_D^2$.
We observe that at large values of $s$ the dependence of
$x_{t/b}$ on $M^2$ is rather small.
While $x_t$ decreases to very small values with
increasing $x_F$, $x_b$ rises towards $x_F$.
Let us review the present knowledge of nuclear effects in deuteron
distribution functions at light-cone momentum fractions probed
in $J/\psi$ production.
Nuclear effects in quark distributions have been studied a great deal
in deep-inelastic scattering processes
(see e.g. \cite{Arneodo94,AnnRev}).
At small values of Bjorken $x$ ($x<0.1$) shadowing effects
in the deuteron of about $(1-4)\%$ were predicted by many models
and also recently observed
(see \cite{NMC94,ShaDeu} and references therein).
They suggest $\epsilon_q (x) \sim (-0.01) \;\mbox{---} \; (-0.04)$ for
$x\sim 0.05 \;\mbox{---} \; 0.001$. At $ x \sim 0.1$ shadowing disappears
and
the deep-inelastic scattering cross section for nuclear targets becomes
slightly larger than the corresponding cross section for free nucleons.
Such a behavior is also expected for deuterium,
although small \cite{FrStLi90}.
It suggests $\epsilon_q (x\sim 0.1) \gsim 0$.
For $x>0.2$ binding effects cause a decreases of
$\epsilon_q <0$, while Fermi motion leads to a rise
at $x>0.8$ (see e.g. \cite{MeScTh94}).
Nuclear effects in gluon distributions are not so well established
up to now. However momentum and baryon number conservation suggest
that for $x<0.2$ they exhibit a behavior similar to
that of the quark distributions --
i.e. $\epsilon_g \sim \epsilon_q$ \cite{FrStLi90}.
The preceeding discussion suggests that modifications of
the deuteron parton distributions might easily be larger than
the charge symmetry violating effects. For example, at a center of
mass energy $\sqrt{s} = 80\,GeV$ and $x_F\approx 0.1$ we have
$x_t \approx 0.02$ and $x_b\approx 0.12$.
While $x_t$ is in the shadowing domain, $x_b$ lies in the
region where the deuteron parton distributions might be enhanced or
are at least equal to the nucleon ones.
The difference
$\Delta \epsilon_{q/g} = \epsilon_{q/g}(x_t) - \epsilon_{q/g}(x_b)$
can therefore easily range from $\Delta \epsilon_{q/g} \approx -0.01$ to
$\Delta \epsilon_{q/g} \approx -0.03$
In contributions to $\Delta\sigma_{pD}$ which are proportional
to charge symmetry violations one may neglect nuclear effects
to a good approximation.
Then nuclear effects enter $\Delta\sigma_{pD}$
via the difference $\Delta \epsilon_{q/g}$ only.
In Fig.~4 we show separately the contributions
of gluon fusion and quark-antiquark annihilation to the ratio
$R_{pD}^{J/\psi} = \frac{\Delta\sigma_{pD}}{d\sigma(J/\psi)_{pp}/dx_F}$,
for different
$\Delta \epsilon_{q/g}$,
at a center of mass energy $\sqrt{s} = 80\,GeV$.
For the charge symmetry violation in the valence distributions we again use
the results of ref.\cite{RoThLo94},
while the charge symmetry violation in the sea and gluon distributions
are chosen to be zero.
We find that at small values of $x_F$ nuclear modifications of the
gluon distribution in deuterium dominate $R_{pD}^{J/\psi}$ or
equivalently $\Delta\sigma_{pD}$.
Therefore, if charge symmetry violation in the gluon distributions,
which are accessible through $R_{pn}^{J/\psi}$, are small,
nuclear modifications of the gluon distribution in deuterium
can be investigated. Figure~4
demonstrates that, at large $x_F$, nuclear effects may easily dominate
charge symmetry violation in the valence distributions.
\section{Conclusion}
We have discussed $J/\psi$ production in proton-neutron and
proton-deuteron collisions as a tool for investigating charge
symmetry breaking and nuclear modifications of parton distributions.
In proton-neutron collisions the difference of the $J/\psi$
production cross section in the forward and backward direction,
$\Delta\sigma_{pn}$, is solely due to charge symmetry violations in the
nucleon.
Charge symmetry breaking in the valence and sea quark distributions
affect the cross section difference significantly only at large
$x_F \sim 0.6$.
At small values of $x_F\lsim 0.1$ a non vanishing result for
$\Delta\sigma_{pn}$ would be entirely due to charge symmetry
violations in the gluon distributions.
Corresponding measurements should be possible
using the neutron beam facility at FNAL.
$J/\psi$ production in the forward and backward direction
through proton-deuteron collisions will be possible at
RHIC for a wide range of center of mass energies and Feynman $x_F$.
The corresponding cross section difference
is, however, not only due to charge symmetry violations but also
to nuclear modifications of the deuteron parton distributions.
As a consequence nuclear modifications of the gluon
distribution in the deuteron could be investigated
as well.
\bigskip
\noindent
We would like to thank J. Moss for helpful discussions.
|
\section{Introduction}
Recently there has been some activity in trying to obtain information
about the structure
of soft Supersymmetry (SUSY)-breaking
terms in effective $N=1$ theories coming from
four-dimensional strings. The basic idea is to identify some $N=1$ chiral
fields
whose
auxiliary components could break SUSY by acquiring a vacuum expectation
value (vev).
No special assumption is made about the possible origin of SUSY-breaking.
Natural
candidates in four-dimensional strings are 1) the complex dilaton field
$S={{4\pi}\over {g^2}}
+ia$ which is present in any four-dimensional string and 2) the moduli fields
$T^i, U^i$ which parametrize the size and shape of the compactified variety
in models obtained by compactification of a ten-dimensional heterotic string.
It is not totally unreasonable to think that some of these fields may play an
important role
in SUSY-breaking. To start with, if string models are to make any sense, these
fields
should be strongly affected by non-perturbative phenomena. They are massless in
perturbation
theory and non-perturbative effects should give them a mass to avoid deviations
from the equivalence principle and other phenomenological problems.
Secondly, these fields are generically present in large classes of
four-dimensional
models (the dilaton in all of them). Finally, the couplings of these fields to
charged matter
are suppressed by powers of the Planck mass, which makes them natural
candidates
to
constitute the SUSY-breaking ``hidden sector'' which is assumed to be present
in phenomenological models of low-energy SUSY.
The important point in this assumption of locating the seed of
SUSY-breaking
in the dilaton/moduli sectors, is that it leads to some interesting
relationships among different soft terms which could perhaps be experimentally
tested.
In ref.\cite{BIM} three of the authors presented a systematic discussion of
the structure of soft terms which may be obtained under the assumption of
dilaton/moduli dominated SUSY breaking in some classes of four-dimensional
strings,
with particular emphasis on the case of Abelian $(0,2)$ orbifold models
\cite{orbifolds}.
We mostly considered a situation in which only the dilaton $S$ and
an ``overall modulus $T$'' field contribute to SUSY-breaking.
In fact, actual four-dimensional strings like orbifolds contain several $T_i$
moduli.
Generic $(0,2)$ orbifold models contain three $T_i$ moduli
fields (only $Z_3$ has 9 and $Z_4$, $Z_6'$ have 5) and a maximum of three
(``complex structure'') $U_i$ fields. The use of an overall modulus $T$ is
equivalent to the assumption that the three $T_i$ fields of generic orbifold
models contribute exactly the same to SUSY-breaking. In the absence
of further dynamical information it is reasonable to expect
similar contributions from the three moduli although not necessarily exactly
the same. In any case it is natural to ask what changes if one
relaxes the overall modulus hypothesis and works with the multimoduli case.
This is one of the purposes of the present paper.
In section 2 we present an analysis of the effects of relaxing
the overall modulus assumption on the results obtained for soft terms.
In the multimoduli case several parameters are needed to specify the
Goldstino direction in the dilaton/moduli space, in contrast with
the overall modulus case where the relevant information is contained
in just one angular parameter $\theta$. The presence of more free
parameters leads to some loss of predictivity for the soft terms.
However, we show that in some cases there are certain
sum-rules among soft terms which hold independently
of the Goldstino direction.
The presence of these sum rules cause that,
{\it on average} the
{\it qualitative} results in ref.\cite{BIM}
still apply. Specifically, if one insists e.g. in
obtaining scalar masses heavier than gauginos
(something not possible at the tree-level
in the approach of ref.\cite{BIM}) ,
this is possible in the multimoduli case, but
the sum-rules often force some of the scalars
to get negative mass$^2$.
If we want to avoid this, we have to
stick to gaugino masses bigger than
(or of order) the scalar masses.
This would lead us back to the qualitative results
obtained in ref.\cite{BIM}.
In the case of
standard model 4-D strings this
tachyonic behaviour may be particularly problematic,
since charge and/or colour could be broken.
In the case of GUTs constructed from strings,
it may just be the signal
of GUT symmetry breaking.
We exemplify the different type of soft terms
which may be obtained in the multimoduli case in some particular
examples, including an $SO(10)$ String-GUT.
Section 3 addresses another simplifying assumption in ref.\cite{BIM}.
There only the case of diagonal kinetic terms for the charged fields
was considered. Indeed this is the generic case in most orbifolds, where
typically some discrete symmetries (or $R$-symmetries) forbid
off-diagonal metrics for the matter fields. On the other hand there are
some orbifolds in which off-diagonal metrics indeed appear and
one expects that in other compactification schemes such metrics may also
appear. This question is not totally academic since, in the presence of
off-diagonal metrics, the soft terms obtained upon SUSY-breaking are
also in general off-diagonal. This may lead to flavour changing
neutral current (FCNC) effects in the low energy effective $N=1$
softly broken Lagrangian.
A third topic of interest is the $B$-parameter, the soft mass term
which is associated to a SUSY mass term $\mu H_1H_2$ for the
pair of Higgsses $H_{1,2}$ in the Minimal Supersymmetric Standard Model (MSSM).
Compared to the other soft
terms, the result for the $B$-parameter is more model-dependent.
Indeed, it depends not only on the dilaton/moduli
dominance assumption but also on the particular mechanism which could
generate the associated ``$\mu$-term''. An interesting possibility to
generate such a term is the one suggested in ref.\cite{GM}
in which it was
pointed out that in the presence of certain bilinear terms in the
K\"ahler potential an effective $\mu$-term of order the gravitino
mass, $m_{3/2}$, is naturally
generated. Interestingly enough, such bilinear terms in the
K\"ahler potential do appear in string models and particularly in
Abelian orbifolds. In section 4 we compute the $\mu $ and
$B$ parameters as well as the soft scalar masses of the
charged fields which could play the role of Higgs particles in
such Abelian orbifold schemes. We find the interesting result that,
independently of the Goldstino direction in the
dilaton/moduli space, one gets the prediction $|tg\beta |=1$
at the string scale. In other words, the direction
$\langle H_1 \rangle =\langle H_2 \rangle $ remains flat
{\it even after SUSY-breaking}.
The results for $B$ corresponding to other sources
for the $\mu$-term are also presented in the multimoduli case
under consideration. In particular, the possibility of generating a small
$\mu$-term from the superpotential \cite{CM} is studied.
We leave some final comments and conclusions for section 5.
\section{ Soft terms: the multimoduli case}
We are going to consider $N=1$ SUSY 4-D strings with
$m$ moduli $T_i$, $i=1,..,m$. Such notation refers to both $T$-type
and $U$-type (K\"ahler class and complex structure in the Calabi-Yau
language) fields.
In addition there will be charged matter fields $C_{\alpha }$ and the
complex dilaton field $S$.
In general we will be considering $(0,2)$ compactifications and thus the
charged
fields do not need to correspond to $27$s of $E_6$.
Before further specifying the class of theories that we are going to consider
a comment about the total number of moduli is in order.
We are used to think of large numbers of $T$ and $U$-like moduli
due to the fact that in $(2,2)$ ($E_6$) compactifications there is a
one to one correspondence between moduli and charged
fields. However, in the case of $(0,2)$ models
with arbitrary gauge group (which is the case of
phenomenological interest) the number of moduli is drastically reduced.
For example,
in the standard $(2,2)$ $Z_3$ orbifold there are 36 moduli $T_i$,
9 associated to the untwisted sector and 27 to the fixed points of the
orbifold.
In the thousands of $(0,2)$ $Z_3$ orbifolds one can construct by adding
different
gauge backgrounds or doing different gauge embeddings, only the
9 untwisted moduli remain in the spectrum. The same applies to models
with $U$-fields. This is also the case for compactifications using
$(2,2)$ minimal superconformal models. Here all singlets associated
to twisted sectors are projected out when proceeding to
$(0,2)$ \cite{Greene}.
So, as these examples
show, in the case of $(0,2)$ compactifications
the number of moduli is drastically reduced to a few fields.
In the case of generic Abelian orbifolds one is in fact left with
only three T-type moduli $T_i$ ($i=1,2,3$), the only exceptions being
$Z_3$, $Z_4$ and $Z'_6$, where such number is 9, 5 and 5 respectively.
The number of $U$-type fields in these $(0,2)$ orbifolds oscillates
between $0$ and $3$, depending on the specific example.
Specifically, $(0,2)$ $Z_2\times Z_2$ orbifolds have 3 $U$ fields,
the orbifolds of type $Z_4,Z_6$,$Z_8,Z_2\times Z_4$,$Z_2\times Z_6$ and
$Z_{12}'$ have just one $U$ field and the rest have no untwisted $U$-fields.
Thus, apart from the three exceptions mentioned above,
this class of models has at most 6 moduli, three of $T$-type (always
present) and at most three of $U$-type. In the case of models obtained
from Calabi-Yau type of compactifications
a similar effect is expected and only one $T$-field associated to the
overall modulus is guaranteed to exist in $(0,2)$ models.
We will consider effective $N=1$ supergravity (SUGRA)
K\"ahler potentials of the
type:
\begin{eqnarray}
& K(S,S^*,T_i,T_i^*,C_{\alpha},C_{\alpha}^*)\ = \
-\log(S+S^*)\ +\ {\hat K}(T_i,T_i^*)\ +\
{\tilde K}_{{\overline{\alpha }}{ \beta }}(T_i,T_i^*){C^*}^{\overline {\alpha}}
C^{\beta }\ & \nonumber\\
&+ \ (\
Z_{{\alpha }{ \beta }}(T_i,T_i^*){C}^{\alpha}
C^{\beta }\ +\ h.c. \ ) \ . &
\label{kahl}
\end{eqnarray}
The first piece is the usual term corresponding to the complex
dilaton $S$ which is present for any compactification whereas the
second is the K\"ahler potential of the moduli fields, where we recall
that we are denoting
the $T$- and $U$-type moduli collectively by $T_i$.
The greek indices label the matter fields and their
kinetic term functions are given by
${\tilde K_{{\overline{\alpha }}{ \beta }}}$ and $Z_{{\alpha }{\beta }}$
to lowest order in the matter fields. The last piece is often forbidden
by gauge invariance in specific models although it may be relevant
in some cases as discussed in section 4.
In this section we are going to consider the case of diagonal metric
both for the moduli and the matter fields and leave the off-diagonal
case for the next section. Then ${\hat K}(T_i,T_i^*)$ will be a sum
of contributions (one for each $T_i$), whereas
${\tilde K_{{\overline{\alpha }}{ \beta }}}$ will be taken of the
diagonal form ${\tilde K_{{\overline{\alpha }}{ \beta }}}
\equiv \delta _{{\overline{\alpha }}{ \beta }} {\tilde K_{\alpha }}$.
The complete $N=1$ SUGRA Lagrangian is determined by
the K\"ahler potential $K({\phi }_M ,\phi^*_M)$, the superpotential
$W({\phi }_M)$ and the gauge kinetic functions
$f_a({\phi }_M)$, where $\phi_M$ generically denotes the chiral fields
$S,T_i,C_{\alpha }$. As is well known, $K$ and $W$ appear in the
Lagrangian only in the combination $G=K+\log|W|^2$. In particular,
the (F-part of the) scalar potential is given by
\begin{equation}
V(\phi _M, \phi ^*_M)\ =\
e^{G} \left( G_M{K}^{M{\bar N}} G_{\bar N}\ -\ 3\right) \ ,
\label{pot}
\end{equation}
where $G_M \equiv \partial_M G \equiv \partial G/ \partial \phi_M$
and $K^{M{\bar N}}$ is the inverse of the K\"ahler metric
$K_{{\bar N }M}\equiv{\partial}_{\bar N}{\partial }_M K$.
The crucial assumption now is to locate the origin of SUSY-breaking in the
dilaton/moduli sector. It is perfectly conceivable that other fields in the
theory, like charged matter fields, could contribute in a leading manner to
SUSY-breaking.
If that is the case, the structure of soft SUSY-breaking terms will be
totally
model-dependent and we would be able to make no model-independent statements at
all
about soft terms. On the contrary, assuming the seed of SUSY-breaking
originates
in the dilaton-moduli sectors will enable us to extract some interesting
results.
We will thus make
that assumption without any further justification.
Let us take the following parametrization
for the vev's of the dilaton and moduli auxiliary
fields $F^S=e^{G/2} G_{ {\bar{S}} S}^{-1} G_{\bar{S}}$ and
$F^i=e^{G/2} G_{ {\bar{i}} i}^{-1} G_{\bar{i}}$:
\begin{equation}
G_{ {\bar{S}} S}^{1/2} F^S\ =\ \sqrt{3}m_{3/2}\sin\theta e^{-i\gamma _S}\ \ ;\ \
G_{ {\bar{i}} i}^{1/2} F^i\ =\ \sqrt{3}m_{3/2}\cos\theta\ e^{-i\gamma _i}
\Theta _i \ \ ,
\label{auxi}
\end{equation}
where $\sum _i \Theta _i^2=1$ and $e^G=m^2_{3/2}$ is the gravitino
mass-squared.
The angle $\theta $ and the $\Theta _i$ just parametrize the
direction of the goldstino in the $S,T_i$ field space.
We have also allowed for the possibility of
some complex phases $\gamma _S, \gamma _i$ which could be relevant
for the CP structure of the theory. This parametrization has the virtue that
when we plug it in the general form of the SUGRA scalar potential
eq.(\ref{pot}), its vev (the cosmological constant) vanishes by
construction. Notice that such a phenomenological approach allows us
to `reabsorb' (or circumvent) our ignorance about the (nonperturbative)
$S$- and $T_i$- dependent part of the superpotential, which is
responsible for SUSY-breaking.
It is now a straightforward
exercise
to compute the bosonic soft SUSY-breaking terms in this class of theories.
Plugging
eqs.(\ref{auxi}) and (\ref{kahl}) into eq.(\ref{pot})
one finds the following results (we recall that we
are considering here a diagonal metric for the matter fields):
\begin{eqnarray}
& m_{\alpha }^2 = \ m_{3/2}^2 \ \left[ 1\ -\ 3\cos^2\theta \
({\hat K}_{ {\overline i} i})^{-1/2} {\Theta }_i e^{i\gamma _i}
(\log{\tilde K}_{\alpha })_{ {\overline i} j}
({\hat K}_{ {\overline j} j})^{-1/2} {\Theta }_j e^{-i\gamma _j} \ \right] \ ,
&
\nonumber \\
& A_{\alpha \beta \gamma } =
\ -\sqrt{3} m_{3/2}\ \left[ e^{-i{\gamma }_S} \sin\theta \right.
& \nonumber \\
& \left. - \ e^{-i{\gamma }_i} \cos\theta \ \Theta_i
({\hat K}_{ {\overline i} i})^{-1/2}
\ \left( {\hat K}_i - \sum_{\delta=\alpha,\beta,\gamma}
(\log {\tilde K}_{\delta })_i
+ (\log h_{\alpha \beta \gamma } )_i \ \right)
\ \right] \ . &
\label{soft}
\end{eqnarray}
The above scalar masses and trilinear scalar couplings correspond
to charged fields which have already been canonically normalized.
Here $h_{\alpha \beta \gamma }$ is a renormalizable
Yukawa coupling involving three charged chiral fields and
$A_{\alpha \beta \gamma }$ is its corresponding trilinear soft term.
Physical gaugino masses $M_a$ for the canonically normalized gaugino fields
are given by $M_a=\frac{1}{2}(Re f_a)^{-1}e^{G/2}{f_a}_M
{K}^{M{\bar N}} G_{\bar N}$.
Since the tree-level gauge kinetic function is given for any 4-D string by
$f_a=k_aS$, where $k_a$ is the Kac-Moody level of the gauge factor,
the result for tree-level gaugino masses is independent of the
moduli sector and is simply given by:
\begin{equation}
M\equiv M_a\ =\ m_{3/2}\sqrt{3} \sin\theta e^{-i\gamma _S} \ .
\label{gaugin}
\end{equation}
As we mentioned above, the parametrization of the auxiliary field vev's
was chosen in such a way to guarantee the automatic vanishing of
the vev of the scalar potential ($V_0=0$). If the value of $V_0$
is not assumed to be zero
the above formulae are modified in the following simple way.
One just has to replace $m_{3/2}\rightarrow Cm_{3/2}$,
where $|C|^2=1+V_0/3m_{3/2}^2$. In addition, the formula for $m_{\alpha }^2$
gets an additional contribution given by $2m_{3/2}^2(|C|^2-1)=2V_0/3$.
The soft term formulae above are in general valid for any compactification
as long we are considering diagonal metrics. In addition one is tacitally
assuming that the tree-level K\"ahler potential and $f_a$-functions
constitute a good aproximation.
The K\"ahler potentials
for the moduli are in general complicated functions.
To illustrate some general features of the multimoduli case
we will concentrate here on the case of generic $(0,2)$ symmetric
Abelian orbifolds. As we mentioned above, this class of models
contains three $T$-type moduli and (at most) three $U$-type moduli.
We will denote them collectively by $T_i$, where e.g. $T_i=U_{i-3}$; $i=4,5,6$.
For this class of models the K\"ahler potential has the form \cite{potential}
\begin{equation}
K(\phi,\phi^*)\ =\ -\log(S+S^*)\ -\ \sum _i \log(T_i+T_i^*)\ +\
\sum _{\alpha } |C_{\alpha }|^2 \Pi_i(T_i+T_i^*)^{n_{\alpha }^i} \ .
\label{orbi}
\end{equation}
Here $n_{\alpha }^i$ are fractional numbers usually called ``modular weights"
of the matter fields $C_{\alpha }$. For each given Abelian orbifold,
independently of the gauge group or particle content, the possible
values of the modular weights are very restricted. For a classification of
modular weights for all Abelian orbifolds see ref.\cite{IL}.
Using the particular form (\ref{orbi}) of the K\"ahler potential and
eqs.(\ref{soft},\ref{gaugin}) we obtain
the following results\footnote{This analysis was also carried out, for the
particular case of the three diagonal moduli $T_i$,
in refs.\cite{japoneses} and \cite{BC},
in order to obtain unification of gauge coupling constants and to analyze
FCNC constraints, respectively.
Some particular multimoduli examples were also considered in
ref.\cite{FKZ} .}
for the scalar masses, gaugino masses and soft trilinear
couplings:
\begin{eqnarray}
&m_{\alpha }^2 = \ m_{3/2}^2(1\ +\ 3\cos^2\theta\ {\vec {n_{\alpha }}}.
{\vec {\Theta ^2}}) \ , &
\nonumber\\
& M = \ \sqrt{3}m_{3/2}\sin\theta e^{-i{\gamma }_S} \ , &
\nonumber\\
& A_{\alpha \beta \gamma } = \ -\sqrt{3} m_{3/2}\ ( \sin\theta e^{-i{\gamma
}_S}
\ +\ \cos\theta \sum _{i=1}^6 e^{-i\gamma _i} {\Theta }^i {\omega
}^i_{\alpha
\beta \gamma } ) \ , &
\label{masorbi}
\end{eqnarray}
where we have defined :
\begin{equation}
{\omega }^i_{\alpha \beta \gamma }\ =\ (1+n^i_{\alpha }+n^i_{\beta
}+n^i_{\gamma
}-
{Y}^i_{\alpha \beta \gamma } )\ \ ;\ \
{Y}^i_{\alpha \beta \gamma } \
= \ {{h^i_{\alpha \beta \gamma }}\over {h_{\alpha \beta \gamma
}}} 2ReT_i \ .
\label{formu}
\end{equation}
Notice that neither the scalar nor the gaugino masses have any explicit
dependence on $S$ or $T_i$, they only depend on the gravitino mass and
the goldstino angles.
This is one of the advantages of a parametrization in terms of such angles.
In the case of the $A$-parameter an explicit $T_i$-dependence may appear in
the term proportional to $Y^i_{\alpha \beta \gamma }$. This explicit
dependence
disappears in three interesting cases: 1) In the dilaton-dominated case
($\cos\theta =0$).
2) When the Yukawa couplings involve only untwisted ({\bf U}) particles,
i.e couplings of the type {\bf UUU}, in which case
the coupling is a constant.
3) When the particles involved in the coupling have all
overall modular weight $n_{\alpha }=-1$ (again, the coupling is
constant). This is possible for couplings of the type
${\bf U}{\bf T}_{-1}{\bf T}_{-1}$, ${\bf T}_{-1}{\bf T}_{-1}{\bf T}_{-1}$,
where the subindex indicates the
value of the overall modular weight of the twisted ({\bf T}) particle
(see below).
This is for example the case of any $Z_2\times Z_2$ orbifold.
There is a fourth case in which the $Y^i_{\alpha \beta \gamma }$-term
does not disappear but is suppressed
for large radii. This happens when the coupling $h_{\alpha \beta \gamma}$
links twisted fields, {\bf TTT}, associated to the same fixed point.
In this case one has $h_{\alpha \beta \gamma}\simeq (constant + O(e^{-T}))$
\cite{FCM} and then $Y^i_{\alpha \beta \gamma }\rightarrow 0$. In all the first
three cases
discussed above the soft terms obtained are independent
of the values of $S$ and $T_i$.
It is appropriate at this point to recall some information about the
``modular weights'' $n_{\alpha }^i$ appearing in these expressions.
For particles belonging to the untwisted sectors one has
\begin{equation}
n_{\alpha }^i \ =\ -\delta ^i_{\alpha }\ ; \ i =1,2,3 ;\ \
n_{\alpha }^i \ =\ -\delta ^{i-3}_{\alpha }\ ; \ i = 4,5,6 \ .
\label{modu}
\end{equation}
Here $i=1,2,3$ labels the three $T$-type moduli and $i=4,5,6$ the three
(maximum) $U$-type moduli, whereas $\alpha =1,2,3$ labels the three
untwisted sectors of the orbifold.
Each twisted sector is associated to an order $N$ twist vector
${\vec v}=(v^1,v^2,v^3)$ defined so that
$0\leq v^i< 1$, $\sum_{i=1}^3v^i=1$. In terms of the $v_i$ one finds
the following modular weights for particles in twisted sectors:
\begin{eqnarray}
& n_{\alpha }^i\ =\ -(1-v^i+p^i-q^i)\ ;\ i=1,2,3 ;\ \ (v^i\not= 0) \ , &
\nonumber\\
& n_{\alpha }^{i+3}\ =\ -(1-v^i+q^i-p^i)\ ;\ i=1,2,3; \ \ (v^i\not=0) \ , &
\nonumber\\
& n_{\alpha }^i=n_{\alpha }^{i+3}\ =\ 0 \ \ (v^i=0) \ , &
\label{modt}
\end{eqnarray}
where $p^i$ and $q^i$ denote the number of (left-handed) oscillator operators
of each chirality in the $i$-th complex direction (see ref.\cite{IL}
for details).
The ``overall T modular weights'' corresponding to the ``overall modulus''
$T$ field considered in ref.\cite{BIM} are given by
$n_{\alpha }=\sum _{i=1}^3 n_{\alpha }^i$.
Twisted sectors with all $v^i\not=0$ (and no oscillators)
have overall modular weights $n_{\alpha }=-2$
due to the property $\sum _{i=1}^3v^i=1$. Twisted sectors with one of the
$v^i$ vanishing have the form ${\vec v}=(1/r,(r-1)/r,0)$
(plus permutations) with $r=2,3,4,6$. Such sectors obviously have
overall modular weights
$n_{\alpha }=-1$. If the twisted particle has also $p$ ($q$)
positive (negative) chirality oscillators, the overall $T$ modular weight
gets an extra addition $=p-q$. Particles with oscillators normally correspond
to small representations of the gauge group (e.g., singlets) so that
one expects the interesting charged particles to be associated
to either untwisted sector or twisted sectors with no oscillators
(or perhaps at most one or two oscillators).
With the above information we can now analyze the different structure of
soft terms available for each Abelian orbifold. The results obtained in
ref.\cite{BIM}
corresponded to the assumption that only $S$ and the overall modulus
$T$ were the seed of SUSY breaking. Within the more general framework here
described, those results correspond to
the particular goldstino direction
\begin{equation}
{\vec {\Theta ^2}}\ =\ ({1\over 3},{1\over 3},{1\over 3},0,0,0)\ \
\label{era}
\end{equation}
and can be recovered from eq.(\ref{masorbi}) and eq.(\ref{formu}) (assuming
also $\gamma_i=\gamma_T$,
$h^i_{\alpha \beta \gamma}=h^T_{\alpha \beta \gamma}/3$):
\begin{eqnarray}
&m_{\alpha }^2 = \ m_{3/2}^2(1\ +\ n_{\alpha}\cos^2\theta) \ , &
\nonumber\\
& M = \ \sqrt{3}m_{3/2}\sin\theta e^{-i{\gamma }_S} \ , &
\nonumber\\
& A_{\alpha \beta \gamma } = \ -\sqrt{3} m_{3/2}\ ( \sin\theta e^{-i{\gamma
}_S}
\ +\ \frac{1}{\sqrt{3}}\cos\theta e^{-i\gamma_T}
{\omega }_{\alpha \beta \gamma } ) \ , &
\label{masorbio}
\end{eqnarray}
where we have defined :
\begin{equation}
{\omega }_{\alpha \beta \gamma }\ =\ (3+n_{\alpha}+n_{\beta}+n_{\gamma}-
{Y}^T_{\alpha \beta \gamma } )\ \ ;\ \
{Y}^T_{\alpha \beta \gamma } \
= \ 2ReT {{h^T_{\alpha \beta \gamma }}\over {h_{\alpha \beta \gamma
}}} \ .
\label{formuo}
\end{equation}
In that case one could extract a number of generic qualitative properties of
soft terms with regard to three important issues : the existence or not
of negative mass$^2$ for some matter fields, the universality of soft
scalar masses,
and the relative sizes of
gaugino versus scalar masses. In the case of an overall $T$ modulus
one finds (see the above formulae):
{\it 1)} Scalars in untwisted and in twisted sectors with overall $T$-modular
weight $n_{\alpha } =-1$
have always masses-squared $\geq 0$.
{\it 2)} Scalars in twisted sectors with $n_{\alpha }\leq -2 $ are always
lighter than those with $n_{\alpha }=-1$. The condition
$\cos^2\theta \leq 1/|n_{\alpha }|$ is required for a particle $C_{\alpha }$
not to become tachyonic.
{\it 3)} Universal soft scalar masses are obtained in two cases: First, in the
dilaton-dominated SUSY-breaking ($\cos\theta =0$) which implies that the
whole soft terms are universal (see eq.(\ref{masorbi}))
\cite{KL,BIM}. Second, if all scalars
have the same overall modular weight $n_{\alpha}=n$ \cite{BIM}.
For example, this always occurs for any $Z_2\times Z_2$ orbifold.
{\it 4)} Due to the above constraints, all scalars $C_{\alpha }$ verify
$M^2\geq m^2_{\alpha }$.
We would like now to study to what extent these general conclusions change
in the multimoduli case. We will discuss them in turn.
{\it 1) Soft masses for $n_{\alpha }=-1 $ particles}
Let us start with the first of these issues, the masses of
$n_{\alpha }=-1$ sectors. There are two types of such sectors, the untwisted
sector (which is present in any orbifold) and the twisted sectors
with $n_{\alpha }=-1$.
We will discuss them in turn. Using the formulae above one finds
the following expressions for scalars in the three untwisted sectors of any
orbifold:
\begin{eqnarray}
& m_1^2\ =\ m_{3/2}^2\ (1-3\cos^2\theta (\Theta ^2_1+\Theta ^2_4) ) \ , &
\nonumber\\
& m_2^2\ =\ m_{3/2}^2\ (1-3\cos^2\theta (\Theta ^2_2+\Theta ^2_5) ) \ , &
\nonumber\\
& m_3^2\ =\ m_{3/2}^2\ (1-3\cos^2\theta (\Theta ^2_3+\Theta ^2_6) ) \ . &
\label{untw}
\end{eqnarray}
One immediately observes that the only way to avoid the presence of
tachyons for {\it any} choice of goldstino direction
in all three sectors is imposing the condition $\cos^2\theta \leq 1/3$.
This is to be compared to the overall modulus case (\ref{masorbio}) in which
positive mass$^2$ was obtained for any $\theta $.
Notice the following important sum-rule which is valid
for the untwisted particles of any orbifold:
\begin{equation}
m_1^2\ +\ m_2^2\ +\ m_3^2\ =\ |M|^2\ \ .
\label{rulix}
\end{equation}
Furthermore, since ${\vec {n_1}}+{\vec {n_2}}+{\vec {n_3}}=-(1,1,1,1,1,1)$
and the {\bf UUU}
Yukawa couplings do not depend on the moduli
one also has
\begin{equation}
A_{123}\ =\ -M \ .
\label{aaa}
\end{equation}
Let us consider now the case of twisted sectors with $n_{\alpha }=-1$.
As we said, the associated twist vectors have the form
${\vec v}=(1/r,(r-1)/r,0)$ (plus permutations) with $r=2,3,4,6$.
Looking at the first of the eqs.(\ref{masorbi})\ one sees that
one has guaranteed a positive mass$^2$ if
$\cos^2\theta \leq r/3(r-1)$. The tighter bound is obtained when $r=6$
which yields $\cos^2\theta \leq 2/5$. A generalization of eqs.(\ref{rulix})
and (\ref{aaa})\ apply also in this case. Consider three particles
$C_{\alpha }$,$C_{\beta }$,$C_{\gamma }$ all with overall modular weight
$=-1$ coupling through a Yukawa $h_{\alpha \beta \gamma }$. They may belong
both
to the untwisted sector or to a twisted sector with $n=-1$, i.e. couplings
of the type ${\bf U}{\bf T}_{-1}{\bf T}_{-1}$,
${\bf T}_{-1}{\bf T}_{-1}{\bf T}_{-1}$.
Then it is easy to
convince oneself that again for any possible twist
${\vec {n_\alpha }}+{\vec {n_\beta }}+{\vec {n_\gamma }}=-(1,1,1,1,1,1)$.
Then one
finds
that for {\it any choice} of goldstino direction
\begin{equation}
m_{\alpha }^2\ +\ m_{\beta }^2\ +\ m_{\gamma }^2\ =\ |M|^2\
=3 m_{3/2}^2\sin^2\theta \
\label{rulox}
\end{equation}
and besides
\begin{equation}
A_{\alpha \beta \gamma }\ =\ -M \ .
\label{ruloxxt}
\end{equation}
The only difference with eqs.(\ref{rulix}), (\ref{aaa}) is that
eqs.(\ref{rulox}), (\ref{ruloxxt}) apply to
any three $n=-1$ particles linked by a Yukawa coupling
(and not only to the three untwisted sectors). Thus, for example, the sum-rule
applies to any set of three particles which couple in any $Z_2\times Z_2$
orbifold.
Specific examples will be shown below.
Notice that if we insist in having a vanishing gaugino mass, the sum-rules
(\ref{rulix}) and (\ref{rulox}) force
the scalars to be either all massless or at least one of them tachyonic.
As we will discuss below, having a tachyonic sector is not necessarily a
problem, it may even be an advantage, so one should not disregard this
possibility
at this point. Of course, in the trivial case when there is no physical
particle in that particular sector which would have negative mass$^2$
the situation is also harmless. Let us show an explicit example of this
possibility.
Consider the second example of Table 3 of ref.\cite{INQ}. This is a
three-generation
$Z_3$
orbifold model with gauge group $SU(3)_c\times SU(3)_L\times SU(3)_R$.
It has the particular property that it has no charged matter in the
untwisted sector so that the sum-rule (\ref{rulix}) can cause no trouble in the
untwisted sector (i.e., no physical tachyons). Consider the goldstino
direction e.g. ${\vec {\Theta }}=(0,0,1)$. The untwisted particles would have
had masses
$m_1^2=m_2^2=m_{3/2}^2$, $m_3^2=m_{3/2}^2(1-3\cos^2\theta )$ whereas the
twisted particles would have
$m_{\bf T}^2=m_{3/2}^2(1-2\cos^2\theta )$. The absence of
charged massless particles in the untwisted sector would have allowed us to
have e.g., $1/3\leq \cos^2\theta \leq 1/2$, values which would have lead to
tachyonic
states in the untwisted sector. For the particular value $\cos^2\theta =1/2$
one gets $m_{\bf T}^2=0$ and gaugino masses $M^2=3/2m^2_{3/2}$.
From the above discussion we conclude that in the
multimoduli case, depending on the goldstino direction, tachyons
may appear both in the untwisted and $n_{\alpha} =-1$ twisted sectors unless
$\cos^2\theta \leq 1/3$. This is to be compared to the overall modulus $T$ case
in which
tachyons never appear. For $\cos^2\theta \geq 1/3 $, one has to
be very careful with the goldstino direction if one is interested
in avoiding tachyons. In some sense, a certain amount of fine tuning
is required so that the goldstino direction goes more and more in the
overall $T$ modulus direction as one increases $\cos^2\theta $.
Nevertheless we should not forget that tachyons, as we already mentioned above,
are not necessarily a problem, but may just show us an instability.
{\it 2) Soft masses for $n_{\alpha }= -2 $ particles }
In the absence of oscillators,
these are particles originated in twisted sectors
${\vec v}=(v^1,v^2,v^3)$ with all $v^i\not=0$. Plugging the expressions for the
modular weights one finds in this case
\begin{equation}
m_{\alpha }^2 =\ m_{3/2}^2(1 - 3\cos^2\theta )\ +\
3m_{3/2}^2\cos^2\theta {\vec v}_{\alpha }.{\vec {\Theta ^2}} \ ,
\label{masorba}
\end{equation}
where ${\vec v}_{\alpha }=(v^1,v^2,v^3,v^1,v^2,v^3)$.
It is obvious from
eq.(\ref{masorba}) that having $\cos^2\theta \leq 1/3$ will be enough to
guarantee the
absence of tachyons for any $n=-2$ particle. This is to be compared with
the overall modulus case analyzed in ref.\cite{BIM}
in which the weaker condition
$\cos^2\theta \leq 1/2$ was required. Notice also that in the overall modulus
$T$ case one always had that the $n=-1$ scalar had bigger masses than the
$n=-2$ scalars. Here the situation may even be reversed.
For any three fields $C_{\alpha}$,$C_{\beta }$,$C_{\gamma}$
linked through a ${\bf T}_{-2}{\bf T}_{-2}{\bf T}_{-2}$ Yukawa coupling
one can check the following sum-rule which is true for any goldstino direction
${\vec {\Theta }}$ :
\begin{equation}
m_{\alpha }^2+m_{\beta }^2+m_{\gamma }^2\ =\ 3m_{3/2}^2(1-2\cos^2\theta)
\ =\ |M|^2\ -\ 3m_{3/2}^2\cos^2\theta \ .
\label{rulax}
\end{equation}
This shows us that, {\it on average}, $n=-2$ twisted particles are lighter than
$n=-1$ particles but the reverse may be true for some particular fields
as long as the above sum-rules are not violated.
It is worth noticing here that twisted Yukawa couplings mixing particles with
$n=-1$ and $n=-2$ are also possible
(e.g. ${\bf T}_{-1}{\bf T}_{-2}{\bf T}_{-2}$,
${\bf T}_{-1}{\bf T}_{-1}{\bf T}_{-2}$). In this case the sum-rule is
\begin{equation}
m_{\alpha }^2+m_{\beta }^2+m_{\gamma }^2\ =\ |M|^2\ -
\ 3m_{3/2}^2\cos^2\theta\ \delta
\label{rulaxxt}
\end{equation}
with
\begin{equation}
\delta\equiv 1- \sum_{k} \Theta_k^2 \ ,
\label{rulaxxtt}
\end{equation}
where $\Theta_k$ are the auxiliary fields of the moduli associated to the
vanishing entry of the $n=-1$ twist vectors (see below eq.(\ref{aaa})) present
in the coupling, i.e. those with $n_{\alpha}^k=0$. Since $0< \delta <1$, the
sum-rule (\ref{rulaxxt}) is rather in-between the (\ref{rulox}) and the
(\ref{rulax}).
Let us finally comment that if the twisted particle has associated an
oscillator
operator, the modular weight decreases in as many units as (positive
chirality) oscillators. This makes very likely for such particles to
have negative mass$^2$ (unless there is approximate dilaton dominance)
. In many cases such particles are just singlets
and such tachyonic behaviour may just denote that these fields are forced
to aquire vev's.
{\it 3) Universality of soft scalar masses}
In the dilaton-dominated case ($\cos\theta =0$) the whole
soft terms are universal
as in the overall modulus case. Also scalars with different overall
modular weights
$n_{\alpha}$ have different masses.
However, unlike the overall modulus case,
non-universal soft scalar masses for particles with the same
$n_{\alpha}$ are allowed and in fact this will be the most general situation
(see e.g. eqs.(\ref{untw},\ref{masorba})).
{\it 4) Gaugino versus scalar masses}
In the overall modulus $T$ discussed in ref.\cite{BIM}
the heaviest scalars were the
ones
with
modular weight $n=-1$ which had mass$^2=|M|^2/3$. So scalars are lighter than
gauginos
at this level. In the multimoduli case sum-rules like (\ref{rulox}) replace
the equation $3m^2_{n=-1}=|M|^2$. In some way, on average the scalars
are lighter than
gauginos but there may be scalars with mass bigger than gauginos. In the case
of
particles with $n=-1$, eq.(\ref{rulox})\ tells us that this can only be true
at
the cost of
having some of the other three scalars with {\it negative} mass$^2$.
This may have diverse phenomenological
implications depending what is the particle content
of the model, as we now explain in some detail:
{\it 4-a) Gaugino versus scalar masses in standard model 4-D strings}
Let us consider first the case of string models with gauge group
$SU(3)_c\times SU(2)_L\times U(1)_Y$$\times G$ and see whether
one can avoid the general situation of ref.\cite{BIM}, where scalar
masses were found to be always smaller than gaugino masses (at tree-level).
In the present more general framework, one can certainly find explicit
examples of orbifold sectors where some individual scalar mass is bigger
than gaugino masses even at the tree-level. For example, let us consider
the case of the $Z_8$ orbifold with an observable particle in the twisted sector
${\bf T}_{\theta^6}$. The modular weight associated to that sector is
${\vec {n_{\theta^6}}}=(-1/4,-3/4,0,0)$ and therefore (see eq.(\ref{masorbi}))
\begin{equation}
m_{\theta^6}^2\ =\ m_{3/2}^2\ \left[1-3\cos^2\theta
\left(\frac{1}{4}\Theta^2_1+\frac{3}{4}\Theta^2_2\right) \right] \ .
\label{masa}
\end{equation}
Then, choosing e.g. a goldstino direction with $\cos^2\theta=5/6$,
$\Theta_1=\Theta_2=0$, one gets $m_{\theta^6}^2=m_{3/2}^2$, $M^2=m_{3/2}^2/2$.
Many more examples along these lines can be found of course. In
general one finds that it is possible to get $m_{\alpha} > M$,
provided $\sin\theta$ is sufficiently small. Indeed, from the general
formulae eq.(\ref{masorbi}) we see that always $m_{\alpha}\leq m_{3/2}$ and
therefore a necessary (although usually not sufficient) condition to get
scalars heavier than gauginos is
\begin{equation}
\cos^2\theta > 2/3 \; .
\label{coseno}
\end{equation}
After such preliminary remark one immediately realizes that,
especially in the case of standard model 4-D strings,
further important restrictions on the possibility of getting
scalars heavier than gauginos come from sum-rules like
(\ref{rulix},\ref{rulox},\ref{rulax},\ref{rulaxxt}),
which typically constrain the masses of three particles linked
via a Yukawa coupling. Suppose that all the three particles involved
are observable particles (squarks, sleptons, Higgses).
If we require that the corresponding squared masses be non-negative
in order to avoid automatically phenomenological problems such
as charge and color breaking or Planck scale Higgs vevs,
then the sum rule will immediately imply that such masses
are smaller than gaugino masses. Conversely, if we tried to
obtain one scalar mass bigger than gaugino masses by an
appropriate choice of the goldstino direction, then at least
one of the other two scalar masses would become tachyonic.
On the other hand, tachyons may be helpful if the particular Yukawa coupling
does not involve observable particles. They could break extra gauge symmetries
and generate large masses for extra particles. We recall that standard-like
models in strings usually have too many extra particles and many extra
U(1) interactions. Although the Fayet-Iliopoulos mechanism helps to cure
the problem \cite{suplemento}, the existence of tachyons is a complementary
solution.
Concerning observable particles, we have just seen that the sum
rules, supplemented by `no-tachyon' requirements, typically lead
to the conclusion that observable scalars are lighter than gauginos
\begin{equation}
m_{\alpha} < M \ ,
\label{masas1}
\end{equation}
similarly to the situation found in the symplified scenario of
ref.\cite{BIM}. Therefore, since gaugino loops play a main
role in the renormalization of scalar masses down to low-energy,
the gluino, slepton and (first and second generation) squark mass
relations at the electroweak scale turn out (again) to be
\begin{equation}
m_l < m_q \simeq M_g \ ,
\label{masas2}
\end{equation}
where gluinos are slightly heavier than squarks. We recall that
slepton masses are smaller than squark masses because they do not
feel the important gluino contribution.
It is still possible to ask whether the generic situation
described by eqs.(\ref{masas1}) and (\ref{masas2}) admits
exceptions.
One possibility
is the following.
One could get some squark or slepton mass bigger than gaugino masses
by allowing a negative soft squared mass for a Higgs field, provided
the total squared Higgs mass (including the $\mu^2$ contribution) is
non-negative\footnote{Notice that such a possibility can be explored in
detail only after specifying the mechanism for generating the $\mu$ parameter
itself (see e.g. ref.\cite{Nuestra}).}.
Another possibility which comes to mind is the case
in which a Yukawa coupling among `observable' particles
originates actually from a non-renormalizable (rather than
renormalizable) coupling\footnote{Notice however that
this is unlikely to be the case for the top Yukawa coupling,
which is relevant e.g. for radiative symmetry breaking.},
where the extra fields in the coupling get vevs (e.g.
$H_2Q_Lu_L^c<\phi...\phi>$ rather than just $H_2Q_Lu_L^c$).
In such a case new sum-rules would apply to the full set of fields in the
coupling and the above three-particle sum-rules could be violated. In
particular, observable scalars would be allowed to be heavier than gauginos,
possibly at the price of having some tachyon among the (standard model
singlet) $\phi$ fields. In both cases
mentioned here one could get a violation of (\ref{masas1}) for some
scalars, i.e.
\begin{equation}
m_{\alpha} > M_a .
\label{masas3}
\end{equation}
However we recall from our initial discussion that this can happen
only for small $\sin\theta$ and special goldstino directions.
Moreover, even for small (but not too small) $\sin\theta$, scalar
and gaugino masses will be still of the same order, so that the
low-energy relation (\ref{masas2}) will still hold.
The only difference is that now squarks,
fulfilling eq.(\ref{masas3}),
will be slightly heavier than
gluinos.
In order to reverse the situation and get instead
\begin{equation}
M_g < m_l , m_q
\label{masas4}
\end{equation}
one needs one of the above `mechanisms' and very small $\sin\theta$,
so that $m_{\alpha} >> M_a$. Note that in such a limit
additional attention should be payed to avoid that a
too large scalar-to-gaugino mass ratio could spoil the
solution to the gauge hierarchy problem.
Before concluding, we recall that a pattern like (\ref{masas4})
for very small $\sin\theta$ was also obtained in the overall modulus
analysis of ref.\cite{BIM} for different reasons, i.e. as an effect of
string loop corrections to $K$ and $f_a$. After the inclusion of
such corrections the masses of gauginos and $n_{\alpha}=-1$ scalars,
which vanish at tree-level for $\sin\theta \rightarrow 0$, become
nonvanishing and typically satisfy relation (\ref{masas3}). One difference
with the previous case is that the loop-induced case gives scalar masses
smaller than $m_{3/2}$ instead than ${\cal O}(m_{3/2})$. In addition,
one may consider this possibility of obtaining scalars heavier than
gauginos as a sort of fine-tuning. In the absence of a more fundamental
theory which tells us in what direction the goldstino angles point, one
would naively say that the most natural possibility would be to assume
that all moduli contribute to SUSY-breaking in more or less (but {\it not}
exactly) the same\footnote{For an explicit example of this, using gaugino
condensation, see ref.\cite{Bailin}.} amount.
Summarizing the situation concerning standard model strings,
we have seen that the overall modulus results are qualitatively
confirmed, in the sense that for generic goldstino directions (with not
too small $\sin\theta$) the low-energy pattern of eq.(\ref{masas2})
typically holds, mainly because of the restrictions coming from
mass sum rules and absence of tachyons. Possible exceptions giving rise
to patterns like (\ref{masas4}) may exist for special goldstino angles,
necessarily including a sufficiently small $\sin\theta$.
{\it 4-b ) Gaugino versus scalar masses in GUT 4-D strings}
What it turned out to be a potential disaster
in the case of standard model strings may be an interesting
advantage in the case of string-GUTs.
In this case it could well be that
the negative mass$^2$ may
just induce gauge symmetry breaking by forcing a vev for a particular
scalar (GUT-Higgs field) in the model.
The latter possibility provides us with interesting phenomenological
consequences.
Here the breaking of SUSY would directly induce further gauge symmetry
breaking.
Let us now show an explicit example of the different possibilities discussed
above (scalars lighter or heavier than gauginos) in the context of GUTS.
We are going to consider a $Z_2\times Z_2$ orbifold
model
which is an $SO(10)$ string-GUT recently constructed in ref.\cite{AFIU}. We
show
in Table 1 the particle content of the model and the quantum numbers of the
particles with respect to the gauge group
$SO(10)\times (SO(8)\times U(1)^2)$. The three untwisted sectors are denoted
by ${\bf U}_1$,${\bf U}_2$,${\bf U}_3$ and the three twisted sectors by
${\bf T}_{\theta}$,${\bf T}_{\omega}$ and ${\bf T}_{\theta \omega}$.
This model has a GUT-Higgs field transforming as a $54$ of $SO(10)$
in the ${\bf U}_3$ untwisted sector. Four net generations as well as
two pairs $16+{\overline {16}}$ are present in the
${\bf T}_{\theta}$,${\bf T}_{\omega}$
twisted sectors. Finally, $10$-plets adequate to do the electro-weak
symmetry breaking belong to the ${\bf T}_{\theta \omega}$ sector.
Yukawa couplings of the following types are present in the
model:
\begin{equation}
{\bf U}_1{\bf U}_2{\bf U}_3\ \ ,\ \
{\bf U}_3{\bf T}_{\theta \omega}{\bf T}_{\theta \omega}\ \ , \ \
{\bf T}_{\theta}{\bf T}_{\omega}{\bf T}_{\theta \omega}
\label{yuk}
\end{equation}
(Not all of the latter two couplings are allowed since the
space-group selection rules may forbid some of them.)
All Yukawa couplings are constants, do not depend on $T_i$
\cite{FCM}.
The $Z_2\times Z_2$ orbifold
has three $T$ moduli and three $U$ moduli in the untwisted sector
but we are considering in this example for simplicity the case in which
only $S$ and the $T_i,i=1,2,3$ participate in SUSY-breaking. The modular
weights
of the
different sectors are:
\begin{eqnarray}
& {\vec {n_1}} =(-1,0,0)\ ;\ {\vec {n_2}}=(0,-1,0)\ ;\ {\vec {n_3}}=(0,0,-1)
\ , &
\nonumber\\
& {\vec {n_{\theta }}} =(0,-1/2,-1/2)\ ;\ {\vec {n_{\omega }}}=(-1/2,0,-1/2)\
;\
{\vec {n_{\theta \omega }}}=(-1/2,-1/2,0) \ . &
\label{modor}
\end{eqnarray}
All the sectors in the $Z_2\times Z_2$ orbifold have overall modular weight
=--1
and hence the sum-rule (\ref{rulox}) applies for any three set of particles
linked by a Yukawa coupling. Notice in particular that
${\vec {n_\alpha }}+{\vec {n_\beta }}+{\vec {n_\gamma }}=-(1,1,1,1,1,1)$
for the sets of particles related by the Yukawas (\ref{yuk}). Thus, for {\it
any goldstino
angle} one has the constraints:
\begin{eqnarray}
& m_1^2+m_2^2+m_3^2\ =\ m_\theta ^2 +m_{\omega }^2+m_{\theta \omega }^2\ =\
m_3^2 +m_{\theta \omega }^2+m_{\theta \omega }^2\ =\ M^2 \ , &
\nonumber\\
& A_{123}\ =\ A_{\theta \omega (\theta \omega )}\ =\
A_{3(\theta \omega )(\theta \omega )}\ =\ -M \ . &
\label{cons}
\end{eqnarray}
To study the different effects of chosing different goldstino directions let us
consider several examples:
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
$Sector $
& $SO(10)\times SO(8)$ & $Q$ & $Q_A$ & $A$ & $
B$ &
$C$ & $D$
\\
\hline
$gauginos $ & $(45,1)+(1,28)$ & 0 & 0
& $3m_{3/2}^2$ & $m_{3/2}^2$ & $m_{3/2}^2$ & $ 0 $
\\
\hline
$ {\bf U}_1 $ & (1,8) & 1/2 & 1/2
& $m_{3/2}^2$ & $0$ & $m_{3/2}^2$ & $m_{3/2}^2$ \\
\hline
& (1,8) & -1/2 & -1/2
& $m_{3/2}^2$ & $0$ & $m_{3/2}^2$ & $m_{3/2}^2$
\\
\hline $ {\bf U}_2 $ &
(1,8) & -1/2 & 1/2 & $m_{3/2}^2$ & $0$ & $m_{3/2}^2$ & $m_{3/2}^2$
\\
\hline & (1,8) &
1/2 & -1/2 & $m_{3/2}^2$ & $0$ & $m_{3/2}^2$ &
$m_{3/2}^2$\\
\hline $ {\bf U}_3 $ &
(54,1) & 0 & 0 & $m_{3/2}^2$ & $m_{3/2}^2$ & $-m_{3/2}^2$ &
$-2m_{3/2}^2$\\
\hline & (1,1) &
0 & 0 & $m_{3/2}^2$& $m_{3/2}^2$ &
$-m_{3/2}^2$ &$-2m_{3/2}^2$\\
\hline
& (1,1) & 0 & 1 & $m_{3/2}^2$ & $m_{3/2}^2$ &$-m_{3/2}^2$ &
$-2m_{3/2}^2$\\
\hline & (1,1)
& 1 & 0 & $m_{3/2}^2$ & $m_{3/2}^2 $ &$-m_{3/2}^2$ &
$-2m_{3/2}^2$\\
\hline & (1,1)
& -1 & 0 & $m_{3/2}^2$ &$m_{3/2}^2 $ & $-m_{3/2}^2$ &
$-2m_{3/2}^2$\\
\hline & (1,1)
& 0 & -1 & $m_{3/2}^2$ & $m_{3/2}^2 $ &$-m_{3/2}^2$ &
$-2m_{3/2}^2$\\
\hline ${\bf T}_{\theta}$
& $3(16,1)$ & 1/4 & 1/4 & $m_{3/2}^2$ & $1/2m_{3/2}^2$ & 0 &
$-1/2m_{3/2}^2$\\
\hline &
$(\overline{16},1)$ & -1/4 & -1/4 & $m_{3/2}^2$ & $1/2m_{3/2}^2$ & 0
&
$-1/2m_{3/2}^2$\\
\hline ${\bf T}_{\omega}$
& $3(16,1)$ & -1/4 & 1/4 & $m_{3/2}^2$ & $1/2m_{3/2}^2$ & 0 &
$-1/2m_{3/2}^2$\\
\hline &
$(\overline{16},1)$ & 1/4 & -1/4 & $m_{3/2}^2$ & $1/2m_{3/2}^2$ & 0
&
$-1/2m_{3/2}^2$\\
\hline
${\bf T}_{\theta \omega}$ & $ 4(10,1)$ & 0 & 1/2 & $m_{3/2}^2$ & $0$ &
$m_{3/2}^2$
&
$m_{3/2}^2$\\
\hline &
$4(10,1)$ & 0 & -1/2 & $m_{3/2}^2$ & $0$ &$m_{3/2}^2$ &
$m_{3/2}^2$\\
\hline &
$3(1,8)$ & 0 & 1/2 & $m_{3/2}^2$ & $0$ &$m_{3/2}^2$ &
$m_{3/2}^2$\\
\hline &
$(1,8)$ & 0 & -1/2 & $m_{3/2}^2$ & $0$ &$m_{3/2}^2$ &
$m_{3/2}^2$\\
\hline &
$8(1,1)$ & 1/2 & 0 & $m_{3/2}^2$ & $0$ &$m_{3/2}^2$ &
$m_{3/2}^2$\\
\hline &
$8(1,1)$ & -1/2 & 0 & $m_{3/2}^2$ & $0$ &$m_{3/2}^2$ &
$m_{3/2}^2$\\
\hline
\end{tabular}
\end{center}
\caption{Particle content and charges of the string-GUT example discussed in
the
text. The four rightmost columns desplay four examples of consistent
soft masses from dilaton/moduli SUSY breaking.}
\label{tuno}
\end{table}
A) Dilaton dominance: $\cos^2\theta =0$. All scalars
have masses $m_{\alpha }^2=m_{3/2}^2$ and
$M^2=3m_{3/2}^2$. The same universal $M/m_{\alpha }$ ratio is
mantained
in the overall modulus case (i.e., ${\vec {\Theta^2 }}=(1/3,1/3,1/3)$)
for any $\theta $. This happens because $n_{\alpha }=-1$ in
eq.(\ref{masorbio}).
B) Consider the goldstino direction ${\vec {\Theta^2 }}=(1/2,1/2,0)$ and
$\cos^2\theta=2/3$. One finds $|M|^2=|A|^2=m_{3/2}^2$ and the scalars get
masses
as shown in column B of Table 1. The soft masses are no longer universal since
e.g. the masses of the electroweak doublets and the generations are
different. This
is important e.g. in computing electro-weak radiative symmetry breaking.
C) Consider the goldstino direction ${\vec {\Theta^2 }}=(0,0,1 )$ and
$\cos^2\theta =2/3$. One still has $|M|^2=|A|^2=m_{3/2}^2$ but now
the GUT-Higgs $54$ and the singlets get negative mass$^2$
(see column C in Table 1). This will
drive a large vev (of order the string scale) $<54>$. Although one would
naively think that the potential becomes unbounded below, one has to recall
that
the matter metrics that we are using are correct to leading order on the matter
fields
and hence for vev's of order of the string scales the potential should be
stabilized.
D) Consider finally the direction ${\vec {\Theta^2 }}=(0,0,1 )$ but
$\cos^2\theta =1$, i.e., only the modulus $T_3$ contributes to SUSY-breaking
(no dilaton contribution). Now the gauginos are massless, the
$10$-plets have positive masses but both the $54$ and the $16+{\overline {16}}$
pairs will tend to get vev's (see column D in Table 1).
As the above examples show, different possibilities are obtained for each given
orbifold model depending on the particular goldstino direction. However,
not any possibility may be realized within a given class of models.
For example, the addition of any combination of soft terms violating
the constraints (\ref{cons}) would be inconsistent with the
hypothesis of dilaton/moduli induced SUSY-breaking. The reader may check that
indeed
the four choices of soft terms shown in the Table verify the constraints in
(\ref{cons}).
Comparing the conclusions of this section with those found in ref.\cite{BIM}
one certainly
finds plenty of differences. However the reader must keep in mind that e.g.
the examples B,C,D above correspond to extreme cases in which some modulus does
not
participate at all in the process of symmetry breaking. On the other hand the
overall modulus case is also in some way an extreme case since the different
moduli
participate in {\it exactly} the same way, which is also a sort of
fine-tuning.
As already mentioned above,
in the absence of a more fundamental theory which tells us in what direction
the
goldstino angles point, one would naively say that the most natural possibility
would be to assume that all moduli contribute to SUSY-breaking in more or
less (but not exactly the same) amount. In this case the conclusions would
be
half-way in-between the results found in this section and those found in
ref.\cite{BIM}.
In this context we must remark the sum-rules discussed above which would be
valid
for any choice of goldstino directions.
Let us finally remark that, in spite of the different possibilities of soft
masses in the multimoduli case, the most natural (slepton-squark-gluino)
mass relations
{\it at low-energy} will be similar to the ones of
the overall modulus case eq.(\ref{masas2}) as shown in point {\it 4-a}.
\section{ Off-diagonal matter metric}
In the previous chapter we confined ourselves to the case of diagonal
matter metric
${\tilde K_{{\overline{\alpha }}{ \beta }}}
\simeq \delta _{{\overline{\alpha }}{ \beta }}$. In fact that assumption is
justified for most of the Abelian orbifold models.
The reason is that, in the case of twisted sectors, each particle has
associated
space-group discrete quantum numbers which forbid off-diagonal metrics
(we are talking here about singular, non-smoothed out $(0,2)$ orbifolds).
In the case of matter fields in untwisted sectors, both gauge invariance and
discrete R-symmetries from the right-moving sector forbids off-diagonal
terms in almost all cases. There are only three exceptions to this general
rule,
the $(0,2)$ models based on the orbifolds $Z_3$,$Z_4$ and $Z_6'$. They are
precisely
the only Abelian orbifolds in which there are more than three $T_i$ moduli,
9, 5 and 5 respectively. They also have in common the existence of an
enhanced non-Abelian gauge symmetry in their $(2,2)$ versions ($SU(3)$ in the
first case, $SU(2)$ in the other two).
An off-diagonal metric only appears for fields in the untwisted
sectors of those examples.
In spite of the relative rareness of
off-diagonal metric in orbifolds, it is worth studying what new features
can appear in this case compared to the diagonal one, since
off-diagonal metrics could be present in other less simple
(e.g., Calabi-Yau) compactifications.
First we go back to eq.(\ref{kahl}) and compute the scalar soft
terms in the most general case where the moduli and matter metrics
are not diagonal.
Then the soft mass matrix ${\cal M}'^2_{ {\overline{\alpha }} { \beta } }$
(corresponding to unnormalized charged fields) and the soft parameters
$A_{\alpha\beta\gamma}$ read
\begin{eqnarray}
\label{mmatrix}
{\cal M}'^2_{{\overline{\alpha }}{ \beta }} & = &
m_{3/2}^2 {\tilde K_{{\overline{\alpha }}{ \beta }}}
- {\overline F}^{\overline{i}} ( \partial_{\overline{i}}\partial_j
{\tilde K_{{\overline{\alpha }}{ \beta }}}
-\partial_{\overline{i}} {\tilde K_{{\overline{\alpha }}{ \gamma}}}
{\tilde K^{{ \gamma} {\overline{\delta}} }}
\partial_j {\tilde K_{{\overline{\delta}}{ \beta}}} ) F^j
\\
A_{\alpha\beta\gamma} & = &
F^S K_S h_{\alpha\beta\gamma} + \delta A_{\alpha\beta\gamma}
\\
\delta A_{\alpha\beta\gamma} & = &
F^i \left[ {\hat K}_i h_{\alpha\beta\gamma}
+ \partial_i h_{\alpha\beta\gamma} - \left(
{\tilde K^{{ \delta} {\overline{\rho}} }}
\partial_i {\tilde K_{{\overline{\rho}}{ \alpha}}} h_{\delta\beta\gamma}
+(\alpha \leftrightarrow \beta)+(\alpha \leftrightarrow \gamma)\right)\right]
\end{eqnarray}
where
\begin{equation}
\label{fgrel1}
F^S = e^{G/2} K_{ {\bar{S}} S}^{-1} G_{\bar S} \; \; , \;\;
F^i = e^{G/2} {\hat K}^{i {\overline j}} G_{\overline j}
\end{equation}
A generalization of the usual `angular parametrization' of
the F-field vev's will be introduced below in a representative example.
The matrix ${\hat K}^{i {\overline j}}$ is the inverse of
the moduli metric
${\hat K}_{ {\overline j} k}=\partial_{\overline j}\partial_k {\hat K}$, i.e.
${\hat K}^{i {\overline j}} {\hat K}_{ {\overline j} k} = \delta^i_k$.
Similarly, for the matter metric, we define
${\tilde K}^{\alpha {\overline \beta}}$ so that
${\tilde K}^{\alpha {\overline \beta}} {\tilde K}_{ {\overline \beta} \gamma}
= \delta^{\alpha}_{\gamma}$. Notice that, after normalizing the fields to get
canonical kinetic terms, the first piece in
eq.(\ref{mmatrix}) will lead to universal diagonal soft
masses but the second piece will generically induce
off-diagonal contributions. Concerning the
$A$-parameters, notice that in this section we have not
factored out the Yukawa couplings as usual, since
proportionality is not guaranteed.
Indeed, although the first term in
$A_{\alpha\beta\gamma}$ is always proportional
in flavour space to the corresponding Yukawa
coupling, the same thing is not necessarily true
for the terms contained in $\delta A$.
One purpose of this section is to study such `off-diagonal' effects
in the soft terms.
In order to get more concrete and manageable results, we will now
particularize the above formulae to
the untwisted sectors of $Z_3$,$Z_4$ and $Z_6'$ orbifolds.
The 9 $T^i$-moduli of the $Z_3$ orbifold enter in the K\"ahler potential as
elements of a $3\times 3$ matrix $T^{\alpha {\overline \beta}}$,
the role of the index $i$ being played by a pair of indices
(with $\alpha, {\overline \beta}=1,2,3$).
Similarly, the 4 $T^i$-moduli of $Z_4$ and $Z_6'$ orbifolds
associated to (say) the first and second complex planes
enter by a $2\times 2$ matrix $T^{\alpha {\overline \beta}}$
(with $\alpha, {\overline \beta}=1,2$). In addition, $Z_4$ ($Z_6'$)
has two additional moduli $T^3$ and $U^3$ (one additional modulus $T^3$)
associated to the third complex plane. Such moduli have diagonal metric,
as well as the associated untwisted fields. On the other side, the
moduli of `matrix' type and the associated untwisted charged fields
have non-diagonal metric, derivable from a K\"ahler potential of
the form
\begin{eqnarray}
\delta K & = & - \log \det\left( (T+T^{\dagger})^{\beta {\overline \alpha}}
- C^{\beta} {\overline C}^{\overline{\alpha}} \right)
\\
& \simeq & - \log \det\ (T+T^{\dagger})^{\beta {\overline \alpha}}
+(T+T^{\dagger})^{-1}_{{\overline \alpha}\beta} {\overline
C}^{\overline{\alpha}}
C^{\beta} \ .
\end{eqnarray}
It is convenient to define the hermitian matrix
\begin{equation}
t \equiv t^{\alpha {\overline \beta}} \equiv
(T+T^{\dagger})^{\alpha {\overline \beta}} \ .
\end{equation}
Then it is easy to find that the metric and inverse metric for moduli
and matter fields have the following simple expressions in terms of $t$:
\begin{equation}
\label{modmetr}
{\hat K}_{ {\overline i} j}=t^{-1}_{{\overline \alpha} \gamma }
t^{-1}_{{\overline \delta} \beta} \;\; , \;\;
{\hat K}^{j {\overline i} }=t^{\gamma {\overline \alpha} }
t^{\beta {\overline \delta}} \;\;
(i\equiv \alpha{\overline \beta} \; , \; j\equiv\gamma{\overline \delta}) \ ,
\end{equation}
\begin{equation}
\label{matmetr}
{\tilde K}_{{\overline \alpha} \beta} = t^{-1}_{{\overline \alpha} \beta}
\;\; , \;\;
{\tilde K}^{\beta {\overline \alpha} } =t^{\beta {\overline \alpha} } \ .
\end{equation}
In addition, the $F^i$'s and $G_i$'s in such sectors are also conveniently
represented by matrices $F \equiv F^{\alpha {\overline \beta}}$ and
$G \equiv \partial{G}/ \partial T^{\alpha {\overline \beta}}$.
The relation between the matrices $F$ and $G$ follows from
eqs.~(\ref{fgrel1}) and (\ref{modmetr}):
\begin{equation}
\label{fgrel2}
F = m_{3/2} t G^* t \ .
\end{equation}
We first consider the $A_{\alpha\beta\gamma}$ parameters,
where the indices can now refer to any untwisted fields of
the orbifolds under study.
The relevant result is that $\delta A_{\alpha\beta\gamma}=0$.
This follows from the above structure of the metric and from the
antisymmetry property of Yukawa couplings with respect to
extra indices (understood above), e.g. $SU(3)$ indices in
(2,2) $Z_3$ orbifolds or $SU(2)$ indices in (2,2) $Z_4$, $Z'_6$
orbifolds. Therefore the result for $A_{\alpha\beta\gamma}$
is simply
\begin{equation}
A_{\alpha\beta\gamma} = F^S K_S h_{\alpha\beta\gamma}
= -\sqrt{3} m_{3/2} \sin\theta e^{-i{\gamma}_S} h_{\alpha\beta\gamma}
\label{ccc}
\end{equation}
which is the same result (after factorizing out the Yukawa coupling as
usual) as for the untwisted sector of
any other orbifold eq.(\ref{aaa}). Thus
{\it even in the presence of off-diagonal metrics
and multiple moduli} the result in eq.(\ref{aaa})
still holds.
We will now consider the soft mass matrix (\ref{mmatrix}) in one
of the sectors with off-diagonal metric. The result can be
written in the following compact form:
\begin{equation}
{\cal M}'^2 =
m_{3/2}^2 t^{-1} - t^{-1} F t^{-1} F^{\dagger} t^{-1} \ .
\end{equation}
If the matter fields are canonically normalized as $C^{\alpha}
\rightarrow {\hat C}^{\alpha} = (t^{-1/2})^{\alpha}_{\beta} C^{\beta}$,
the normalized soft mass matrix can be written as
\begin{equation}
\label{mdelta}
{\cal M}^2 = m_{3/2}^2 ( 1 - \Delta) \ ,
\end{equation}
where $1$ stands for the unit matrix and the $\Delta$ is the matrix
\begin{equation}
\label{delmat}
\Delta = \frac{1}{m_{3/2}^2} t^{-1/2} F t^{-1} F^{\dagger} t^{-1/2} \ .
\end{equation}
It is interesting to notice that the contribution to SUSY-breaking
from the moduli of such a sector is
\begin{equation}
{\overline F}^{\overline i} {\hat K}_{ {\overline i} j } F^j
= m_{3/2}^2 {\mbox {Tr}} \Delta \ .
\end{equation}
To continue the discussion we will focus for definiteness
on the case of $Z_3$, where the 9 moduli $T^{\alpha {\overline \beta}}$
exhaust the set of untwisted moduli. We can consider the following
parametrization of the dilaton/moduli SUSY-breaking:
\begin{equation}
(S+S^*)^{-1} F^S = \sqrt{3} m_{3/2} \sin\theta e^{-i \gamma_S}
\;\; ; \;\;
t^{-1/2} F t^{-1/2} = \sqrt{3} m_{3/2} \cos\theta \Theta \ ,
\end{equation}
where $\Theta$ is a $3\times 3 $ matrix satisfying
\begin{equation}
{\mbox {Tr}} \Theta \Theta^{\dagger} = 1 \ .
\end{equation}
Notice that the matrix $\Delta$ in ${\cal M}^2$ (\ref{mdelta})
can be written
\begin{equation}
\label{deldel}
\Delta = 3 \cos^2\theta \, \Theta \Theta^{\dagger} \ .
\end{equation}
In particular, from this one immediately sees that: 1) $\Delta$ is
positive definite and ${\mbox {Tr}}\Delta=3 \cos^2\theta$ ;
2) the sum of the three eigenvalues of ${\cal M}^2$ satisfies
\begin{equation}
{\mbox {Tr}} {\cal M}^2 = 3 m_{3/2}^2 \sin^2\theta = |M|^2
\label{ddd}
\end{equation}
which confirms the already stated sum-rule eq.(\ref{rulix}) for untwisted
matter
in orbifolds, {\it even in the presence of off-diagonal metrics}.
An interesting question related to flavour changing issues\footnote{These
were analyzed for the simplest case of diagonal metric in
refs.\cite{BIM,LN}.}
concerns the degree of degeneracy among the three eigenvalues of ${\cal M}^2$.
It is clear that, for generic values (vev's) of the matrices $t$ and $F$
(or $\Theta$), $\Delta$ will have a generic matrix structure
and therefore the eigenvalues of ${\cal M}^2$ will be non-degenerate.
The approximately degenerate case occurs only when ${\cal M}^2$ is
approximately proportional to the unit matrix\footnote{This corresponds
to the simplest way of avoiding FCNC. Another possibility occurs
if scalar and fermionic mass matrices happen to be aligned \cite{NS}.
This and other issues on FCNC would require a detailed analysis of the flavour
structure of the models, which go beyond the scope of the present paper.},
i.e. ${\cal M}^2 \propto 1$.
This happens: 1) when $\Delta \ll 1$ ; 2) when $\Delta \propto 1$.
1) $\Delta \ll 1$. This happens when $\cos^2\theta \ll 1$, i.e.
when the contribution of the moduli $T^{\alpha {\overline \beta}}$
to SUSY-breaking is negligible. In the case of $Z_3$ this
just corresponds to the dilaton dominated SUSY-breaking
(in the case of $Z_4, Z_6'$ the SUSY-breaking could be shared between
$S$ and the third-complex-plane moduli). Actually, when discussing
FCNC constraints on soft masses, one should consider the
renormalization effects from the string scale to the electroweak
scale. Such effects include flavour independent contributions from
gauginos. For example, if squarks originated from a sector like
the one under study, the low energy mass matrix would read
${\cal M}^2(M_Z) \sim m_{3/2}^2 ( (1+ 24 \sin^2\theta) 1 - \Delta)$,
with $\Delta$ as in eq.(\ref{deldel}) for $Z_3$.
Then the constraint $\cos^2\theta \ll 1$
would be relaxed to $\cos^2\theta \ll 1+ 24 \sin^2\theta$ \cite{BIM}
and the moduli would be allowed to participate to some extent to
SUSY-breaking. On the other side, no significant relaxation would be
obtained for sleptons.
2) $\Delta \propto 1$. This condition guarantees that ${\cal M}^2 \propto 1$
even when the moduli participate significantly to SUSY-breaking. Observing
eq.(\ref{delmat}), we can distinguish two subcases.
2a) If $t$ and $F$ are treated as independent objects, than the only
obvious way to satisy that condition is that both $t \propto 1$ and
$F \propto 1$. This requires not only that the off-diagonal moduli
and F-terms be negligible, but also that the diagonal ones be almost
identical, i.e. one is pushed towards the overall modulus limit.
2b) Such conclusion may be evaded if $t$ and $F$ are related
in some way, e.g. if $F \propto t$ (giving again $\Delta \propto 1$).
If this were the case, the off-diagonal elements of $F$ and $t$ would
not need to be negligible with respect to the diagonal ones.
An extreme example of this situation happens when $W$ does not depend
on the $T^{\alpha {\overline \beta}}$. In that case $F = -m_{3/2} t$
and $\Delta = 1$, implying ${\cal M}^2=0$ and a no-scale scenario.
An example where ${\cal M}^2 \neq 0$ can be obtained e.g. if $W$ depends
on $T^{\alpha {\overline \beta}}$ only via $\det T^{\alpha {\overline \beta}}$
(and if the vev of $T^{\alpha {\overline \beta}}$ is hermitian).
\section{The B-parameter and the $\mu $ problem}
When an (effective) $N=1$ SUSY mass
$\mu _{\alpha \beta }C^{\alpha }C^{\beta }$
appears in the Lagrangian of an $N=1$ theory,
SUSY-breaking also induces an associated
SUSY-breaking term
$B_{\alpha \beta }\mu_{\alpha \beta } C^{\alpha }C^{\beta }+h.c.$.
Very often these terms are absent due to gauge invariance. Thus in the
MSSM there is only one $B$-term associated to a
possible $\mu H_1H_2$ SUSY mass term. In fact both a $\mu$-term and a
$B$-term are phenomenologically required in the MSSM in order to, among other
things, avoid the presence of a visible axion.
The parameter $\mu $ of the MSSM has to be (on phenomenological grounds) of the
order of the low-energy SUSY-breaking scale (i.e., of order $m_{3/2}$).
The absence of a symmetry reason for such small value for $\mu $ is called
the ``$\mu $-problem" \cite{review}. Thus in order to be able to compute
$B$-term in a given model, we need first a mechanism which might naturally
induce a $\mu $-term of order $m_{3/2}$. We will discuss some of the
mechanisms proposed within the context of string-models to solve this
$\mu$-problem and we will also provide expressions for the associated
$B$-terms in this section.
\subsection{$B$-term from the K\"ahler potential in orbifold models}
It was pointed out in ref.\cite{GM} that terms in a K\"ahler potential
like the one proportional to $Z_{\alpha \beta }$ in eq.(\ref{kahl})
can naturally induce a $\mu $-term for the $C_{\alpha }$ fields
of order $m_{3/2}$ after SUSY-breaking, thus providing a rationale
for the size of $\mu $. Recently it has been realized that such type of
terms do appear in the K\"ahler potential of some Calabi-Yau type
compactifications \cite{KL}
and in orbifold models \cite{LLM,AGNT,FKZ}. Let us consider the case in which
e.g., due to gauge invariance, there is only one possible $\mu $-term
(and correspondingly one $B$-term) associated to a pair of matter fields
$C_1$,$C_2$. From eqs.(\ref{kahl},\ref{pot},\ref{auxi})
and from the fermionic part of the SUGRA lagrangian
one can check that a SUSY mass term $\mu C_1 C_2$
and a scalar term $B \mu (C_1 C_2) +h.c.$
are induced upon SUSY-breaking in the effective low energy theory
(here the kinetic terms for $C_{1,2}$ have been normalized to one).
If we introduce the abbreviations
\begin{equation}
L^Z \equiv \log Z \;\; , \;\;
L^{\alpha} \equiv \log {\tilde K}_{\alpha } \;\; , \;\;
X \equiv 1 - \sqrt{3} \cos\theta \ e^{i\gamma _i}{\Theta _i}
({\hat K}_{ {\overline i} i})^{-1/2} L_{\overline i}^Z
\label{xxx}
\end{equation}
the $\mu$ and $B$ parameters (we will call them $\mu_Z$ and $B_Z$) are given by
\begin{eqnarray}
& \mu_Z \ =\ m_{3/2} ( {\tilde K}_1 {\tilde K}_2 )^{-1/2} Z X \ , &
\label{mmu}
\\
& B_Z\ =\ m_{3/2} X^{-1}
\left[ 2 + \sqrt{3} \cos\theta ({\hat K}_{ {\overline i} i})^{-1/2}
{\Theta_i }
\left( e^{-i\gamma _i}
( L_i^Z - L^1_i - L^2_i )
-e^{i\gamma _i} L_{\overline i}^Z \right) \ \right.
& \nonumber\\
& \left. +
\ 3 \cos^2\theta ({\hat K}_{ {\overline i} i})^{-1/2}
{\Theta_i } e^{i\gamma _i} \ \left(
L_{\overline i}^Z ( L^1_j+L^2_j)
- L_{\overline i}^Z L_j^Z - L_{{\overline i} j}^Z\ \right)
({\hat K}_{ {\overline j} j})^{-1/2}
{\Theta _j } e^{-i\gamma _j} \right] \ . &
\label{bcy}
\end{eqnarray}
The above formulae apply to the cases where the moduli on
which ${\tilde K}_1(T_i,T_i^*)$, ${\tilde K}_2(T_i,T_i^*)$ and
$Z(T_i,T_i^*)$ depend
have diagonal metric, which is the relevant case we are
going to discuss (anyway, the above formulae are easily generalized
to more general situations).
If the value of $V_0$ is not assumed to be zero, one just has to replace
$\cos\theta \rightarrow C\cos\theta$ in eqs.(\ref{xxx},\ref{mmu},\ref{bcy}),
where $C$ is given below eq.(\ref{gaugin}).
In addition, the formula for $B$ gets an additional contribution
given by $m_{3/2} X^{-1} 3(C^2-1)$.
It has been recently shown that the untwisted sector of orbifolds
with at least one complex-structure field $U$ possesses the required
structure $Z(T_i,T_i^*)C_1C_2+h.c.$ in their K\"ahler
potentials. Specifically, the $Z_N$ orbifolds
based on $Z_4,Z_6$,$Z_8,Z_{12}'$ and the $Z_N\times Z_M$ orbifolds based
on $Z_2\times Z_4$ and $Z_2\times Z_6$ do all have a $U$-type field in (say)
the third complex plane. In addition the $Z_2\times Z_2$ orbifold has $U$
fields in the three complex planes.
In all these models the piece of the K\"ahler potential involving
the moduli and the untwisted matter fields $C_{1,2}$ in the third complex
plane has the form
\begin{eqnarray}
& K(T_i,T_i^*,C_1,C_2)=K'(T_l,T_l^*)
& \nonumber\\
& -\log\left((T_3+T_3^*)(U_3+U_3^*) - (C_1+C_2^*)(C_1^*+C_2)\right)
& \label{kahlb} \\
& \simeq
K'(T_l,T_l^*)
- \log(T_3+T_3^*) - \log(U_3+U_3^*)\ +
\frac{(C_1+C_2^*)(C_1^*+C_2)}{(T_3+T_3^*)(U_3+U_3^*)}
\label{kahlexp}
\end{eqnarray}
The first term $K'(T_l,T_l^*)$ determines the (not necessarily diagonal)
metric of the moduli $T_l \neq T_3, U_3$ associated to the first and
second complex planes. The last term describes an
$SO(2,n)/SO(2)\times SO(n)$ K\"ahler manifold ($n=4$ if we
focus on just one component of $C_1$ and $C_2$) parametrized by
$T_3, U_3, C_1, C_2$. If the expansion shown in (\ref{kahlexp}) is
performed, on one hand one recovers the well known
factorization $SO(2,2)/SO(2)\times SO(2) \simeq (SU(1,1)/U(1))^2$
for the submanifold spanned by $T_3$ and $U_3$ (which have
therefore diagonal metric to lowest order in the matter fields),
whereas on the other hand one can easily identify the
functions $Z, {\tilde K}_1, {\tilde K}_2$ associated to $C_1$ and $C_2$:
\begin{equation}
Z\ =\ {\tilde K}_1 \ =\ {\tilde K}_2\ =\ {1\over {(T_3+T_3^*)(U_3+U_3^*)}}
\ .
\label{zzz}
\end{equation}
Plugging back these expressions in eqs.(\ref{mmu},\ref{bcy},\ref{xxx})
one can easily compute $\mu$ and $B$ for this interesting class
of models:
\begin{eqnarray}
& \mu_Z \ =\ m_{3/2}\ \left( 1\ +\ \sqrt{3}\cos\theta
(e^{i \gamma_3} \Theta _3 + e^{i \gamma_6} \Theta _6)\right) \ , &
\label{muu}
\\
& B_Z\mu_Z=2m_{3/2}^2 \left( 1+\sqrt{3} \cos\theta
( \cos\gamma_3 \Theta_3 + \cos\gamma_6 \Theta_6) \ \right.
& \nonumber\\
& \left.
+\ 3\cos^2\theta \cos(\gamma_3-\gamma_6) {\Theta _3}{\Theta _6} \right) \ . &
\label{bmu}
\end{eqnarray}
In addition, we recall from eq.(\ref{untw}) that the soft masses are
\begin{equation}
m^2_{C_1}\ =\ m^2_{C_2}\ =\ m_{3/2}^2\ \left( 1\ -\ 3\cos^2\theta
(\Theta_3^2+\Theta _6^2)\right) \ .
\label{mundos}
\end{equation}
In general, the dimension-two scalar potential for $C_{1,2}$
(now denoting again normalized fields) after SUSY-breaking has
the form
\begin{equation}
V_2(C_1,C_2)\ =\ (m_{C_1}^2+|\mu|^2)|C_1|^2 + (m_{C_2}^2+|\mu| ^2)|C_2|^2
+(B\mu C_1C_2+h.c.)\
\label{flaty}
\end{equation}
In the specific case under consideration, from
eqs.(\ref{muu},\ref{bmu},\ref{mundos}) we find the remarkable result,
which is also true for any value of $C$,
that the three coefficients
in $V_2(C_1,C_2)$ are equal, i.e.
\begin{equation}
m_{C_1}^2+|\mu_Z |^2 = m_{C_2}^2+|\mu_Z| ^2 = B_Z\mu_Z
\label{result}
\end{equation}
so that $V_2(C_1,C_2)$ has the simple form
\begin{equation}
V_2(C_1,C_2)\ =\ B_Z\mu_Z \ (C_1+C_2^*)(C_1^*+C_2) \ .
\label{potflat}
\end{equation}
Although the common value of the three coefficients in eq.(\ref{result})
depends on the Goldstino direction via the parameters
$\cos\theta$, $\Theta_3$, $\Theta_6$,\ldots (see expression of $B_Z\mu_Z$
in eq.(\ref{bmu})), we stress that the equality itself and the form
of $V_2$ hold {\em independently of the Goldstino direction}.
The only constraint that one may want to impose is that the coefficient
$B_Z\mu_Z$ be non-negative, which would select a region of parameter space.
For instance, if one neglects phases, such
requirement can be written simply as
\begin{equation}
(1+\sqrt{3} \cos\theta \ \Theta_3) (1+\sqrt{3} \cos\theta \ \Theta_6) \geq 0
\ .
\end{equation}
We notice in passing that the fields $C_{1,2}$ appear in the
SUSY-breaking scalar potential in the same combination as in
the K\"ahler potential. This particular form may be understood as due to
a symmetry under which $C_{1,2}\rightarrow C_{1,2}+i\delta $ in the K\"ahler
potential which is transmitted to the final form of the scalar potential.
An important (Goldstino-direction-independent) consequence of the above
form (\ref{potflat}) is that $V_2(C_1,C_2)$ identically vanishes
along the direction $C_1 = - C_2^*$, on which gauge symmetry is broken.
If dimension-four couplings respect such flat direction (which is
certainly the case for D-terms), we arrive at the important result
that along $<C_1>=-<C_2^*>$ the {\it flatness is not
spoiled by the dilaton/moduli induced SUSY-breaking}.
This is certainly a very remarkable property.
This result can be rephrased in terms of the usual parameter
$\tan\beta=<C_2>/<C_1>$ (we now assume real vev's).
It is well known that, for a potential of the generic form
(\ref{flaty}) (+D-terms), the minimization conditions yield
\begin{equation}
\sin2\beta \ =\ { {-2 B\mu} \over {m_{C_1}^2+m_{C_2}^2+2|\mu|^2} } \ .
\label{sbet}
\end{equation}
In particular, this relation embodies the boundedness requirement:
if the absolute value of the right-hand side becomes bigger than one,
this would indicate that the potential becomes unbounded from below.
As we have seen, in the class of models under consideration
the particular expressions of the mass parameters lead to
the equality (\ref{result}), which in turns implies
$\sin 2\beta= -1$. Thus one finds $\tan\beta =-1$
{\it for any value of $\cos\theta $,$\Theta _3 $,$\Theta _6 $} (and of
the other $\Theta_i$'s of course), i.e. for any Goldstino direction.
It is interesting to relate these results to similar ones
obtained in ref.\cite{BZ} in a slightly different context. In
ref.\cite{BZ} a {\em specific} SUGRA model was built,
where the Higgs-dependent part of the K\"ahler potential
had the form in eq.(\ref{kahlb}), with
$T_3=U_3$. The geometrical properties of the associated
manifold and a simple choice for the superpotential allowed
to obtain the simultaneous breaking of SUSY
and gauge symmetry, with the cosmological constant identically
vanishing along some flat directions which included the
$|C_1| = |C_2|$ direction. This also implied a partial participation
of charged fields in the process of SUSY-breaking\footnote{An elaboration
of this idea was later studied in ref.\cite{BFZ}.}. In the
limit of suppressed goldstino components along the Higgsinos,
SUSY-breaking was essentially dilaton/moduli dominated.
Then such model could be viewed as a very special
case of the more general framework here discussed,
characterized by specific values of the goldstino angles:
$\cos^2\theta=2/3$, $\Theta_3^2=\Theta_6^2=1/2$ and vanishing
values for the remaining $\Theta_i$'s. In particular
one had $V_2(C_1,C_2)\equiv 0$, the flat direction
$|C_1|=|C_2|$ being enforced by the D-term.
The remarkable result obtained in this section is that
the prediction $|\tan\beta|=1$ is actually valid for
a much broader class of models and holds irrespectively
of the goldstino direction in the dilaton/moduli space.
Whether the above mechanism can be successfully implemented
in the case of the electroweak Higgs fields remains
an open question. Flat potentials of the type here considered
could be interesting also for the breaking of a grand-unified
gauge group (as suggested e.g. in ref.\cite{BFZ}), in particular in
the context of models like string-GUTs \cite{AFIU}, in which
a vev of order the string scale is not problematic.
As an additional comment, it is worth recalling that in previous
analyses of the above mechanism for generating $\mu$ and $B$
in the string context \cite{KL,BLM,BIM} the value of $\mu$ was left
as a free parameter since one did not have an explicit expression for
the function $Z$. However, if the explicit orbifold formulae for
$Z$ are used, one is able to predict both $\mu$ and $B$ reaching
the above conclusion. We should add that situations are conceivable
where the above result may be evaded, for example if the physical Higgs
doublets are a mixture of the above fields with some other doublets coming
from other sectors (e.g. twisted) of the theory.
\subsection{$B$-term from the superpotential}
There is an alternative mechanism to the one studied in the previous subsection
to generate a $B$-term in the scalar potential.
It is well known that if the superpotential $W$ is assumed to have a
$\mu C_1C_2$ SUSY mass term, $\mu$ being an initial parameter, then a
$B$-term is automatically generated. We will call it $B_{\mu}$.
If we introduce the abbreviation
\begin{equation}
L^{\mu} \equiv \log {\mu}
\label{xxxx}
\end{equation}
the $\mu$ and $B$ parameters are given by
\begin{eqnarray}
& {\mu'} \ =\ {\mu} e^{K/2} \frac{W^*}{|W|}
({\tilde K}_1 {\tilde K}_2 )^{-1/2} \ , &
\label{mmuu}
\\
& B_{\mu}\ =\ m_{3/2}
\left[ -1 - \sqrt{3} e^{-i\gamma_S} \sin\theta (1- L_S^{\mu} 2ReS)
\right.
& \nonumber\\
& \left. +
\sqrt{3} \cos\theta ({\hat K}_{ {\overline i} i})^{-1/2}
{\Theta_i } e^{-i\gamma _i}
({\hat K}_i + L_i^{\mu} - L^1_i - L^2_i ) \right] \ , &
\label{bcyy}
\end{eqnarray}
where the low-energy SUSY mass ${\mu'}$ is related to ${\mu}$ via
the usual SUGRA rescaling,
and again the kinetic terms for $C_{1,2}$ have been normalized to one. In
the above formulae we have assumed that in general ${\mu}$ will depend on
the SUSY-breaking sector fields, i.e. ${\mu}={\mu}(S,T_i)$.
These formulae are completely general and valid for any solution to the
${\mu}$-problem which introduces a small mass term $\mu(S,T_i) C_1C_2$ in $W$.
This type of solutions exists.
In ref.\cite{CM} was pointed out that the presence of a non-renormalizable
term in the superpotential
\begin{equation}
\lambda W C_1 C_2
\label{norenor}
\end{equation}
characterized by the coupling $\lambda$, yields dynamically a ${\mu}$ parameter
when $W$ acquires a vev
\begin{equation}
\mu = \lambda W \ .
\label{vev}
\end{equation}
The fact that $\mu$ is small is a consequence of our assumption of a correct
SUSY-breaking scale $m_{3/2}=e^{G/2}=e^{K/2}|W|$. The superpotential
eq.(\ref{norenor}) which provides a possible solution to the $\mu$ problem
can naturally be obtained in the context of strings.
A realistic example where non-perturbative SUSY-breaking mechanisms like
gaugino-squark condensation induce that superpotential was given in
ref.\cite{CM},
where $\lambda=\lambda(T_i)$ is a non-renormalizable Yukawa coupling between
the Higgses and the squarks and after eliminating the gaugino and squarks
bound states $W=W(S,T_i)$.
In ref.\cite{AGNT} the same kind of superpotential was obtained through
pure gaugino condensation in orbifolds with at least one complex-structure
field $U$. This is because in these orbifolds
matter field-dependent threshold corrections
($\propto C_1C_2$) appear in the gauge kinetic function $f$. We recall
that after eliminating the gaugino bound states the non-perturbative
superpotential $W\sim exp(3f/2b_0)$, where $b_0$ is the one-loop
$\beta$-function coefficient of the ``hidden'' gauge group.
After expanding the exponential, the superpotential will have a
contribution of the type (\ref{norenor}).
Again, $\lambda=\lambda(T_i)$, since the above
proportionality factor due to threshold corrections
depends on Dedekind functions which depend in turn on the moduli.
So with this solution (\ref{vev}) to the $\mu$-problem in strings:
\begin{equation}
\mu(S,T_i) = \lambda(T_i) W(S,T_i) \ .
\label{strings}
\end{equation}
Plugging back this expression in eqs.(\ref{mmuu},\ref{bcyy}) and imposing
the vanishing of the cosmological constant $V_0$, one can
easily compute $\mu$ and $B$ for this mechanism. We will call them
$\mu_{\lambda}$ and $B_{\lambda}$
\begin{eqnarray}
& \mu_{\lambda} \ =\ {\lambda} m_{3/2}
({\tilde K}_1 {\tilde K}_2 )^{-1/2} \ , &
\label{mmuuu}
\\
& B_{\lambda}\ =\ m_{3/2}
\left[ 2 + \sqrt{3} \cos\theta ({\hat K}_{ {\overline i} i})^{-1/2}
{\Theta_i } e^{-i\gamma _i}
(L_i^{\lambda} - L^1_i - L^2_i ) \right] \ , &
\label{blambda}
\end{eqnarray}
where
\begin{equation}
L^{\lambda} \equiv \log {\lambda} \ .
\label{xxxxx}
\end{equation}
If the value of $V_0$ is not assumed to be zero, one just has to replace
$\cos\theta \rightarrow C\cos\theta$ and
$\sin\theta \rightarrow C\sin\theta$
in eqs.(\ref{bcyy},\ref{blambda}),
where $C$ is given below eq.(\ref{gaugin}).
In addition, the formula for $B_{\lambda}$, eq.(\ref{blambda}),
gets an additional contribution
given by $m_{3/2} 3(C^2-1)$.
Concentrating again on the interesting case of orbifolds, where the
K\"ahler potential eq.(\ref{orbi}) is known, we obtain from eq.(\ref{blambda})
\begin{eqnarray}
& B_{\lambda}\ =\ m_{3/2}
\left[ 2 - \sqrt{3} \cos\theta \sum _{i=1}^6 e^{-i\gamma _i} {\Theta }_i
\left(n_1^i + n_2^i - \frac{\lambda_i}{\lambda} 2 Re T_i\right) \right] \ . &
\label{bcyyy}
\end{eqnarray}
Notice that it is conceivable that both mechanisms, the one solving the
$\mu$-problem through the K\"ahler potential (see subsection 4.1) \cite{GM}
and the other one solving it through the superpotential \cite{CM} shown above,
could be present simultaneously. In that case the general expressions for
$B$ and $\mu$ are easily obtained
\begin{eqnarray}
& \mu \ =\ \mu_Z + \mu_{\lambda} \ , &
\label{mmuuuuu}
\\
& B\ =\ \mu^{-1} (B_Z \mu_Z + B_{\lambda} \mu_{\lambda}) \ , &
\label{bcyyyy}
\end{eqnarray}
where $\mu_Z$, $B_Z$ are given in eqs.(\ref{muu},\ref{bmu}).
For example, in the case of orbifolds with at least one complex-structure
field $U$, where the $B_Z$-term from the K\"ahler potential is present, if
a gaugino condensate is formed, then automatically the $B_{\lambda}$-term
from the superpotential is also present as mentioned above. Now, as in the case
of $B_Z$ (see eqs.(\ref{muu},\ref{bmu})),
in $B_{\lambda}$ (\ref{bcyyy}) only ${\Theta}_3$ and ${\Theta }_6$
contribute. We recall that
the values of ${\tilde K}_1, {\tilde K}_2$ are given by eq.(\ref{zzz})
and besides, $\lambda=\lambda(T_3,U_3)$ (the concrete expression can be found
in ref.\cite{AGNT}). However, in this case the last equality of
eq.(\ref{result}) with $Z \rightarrow \lambda$ does not hold.
\section{Final comments and conclusions}
In this paper we have generalized in several directions
previous analyses of SUSY-breaking soft terms
induced by dilaton/moduli sectors. In particular, we
have studied the new features appearing when one goes to the
Abelian orbifold multimoduli case. We have found that there are qualitative
changes in the general patterns of soft terms. In some way
({\it on average}) the results are similar to the case in which
only $S$ and the ``overall modulus'' $T$ field are considered.
However, if one examines the soft terms for each particle individually
one finds different extreme patterns.
For example, non-universal soft scalar masses for particles with the
same overall modular weight are allowed and in fact this will be the
most general situation.
Besides, unlike in the
case considered in \cite{BIM}, gauginos may be lighter than
scalars even at the tree-level.
The possibilities are, however, not arbitrary. The fact that
{\it on average}
the results are similar to the simple $S,T$ case are embodied in
general sum rules like those in
eqs.(\ref{rulix},\ref{rulox},\ref{rulax},\ref{rulaxxt})
which relate soft terms of different particles in the theory.
Due to the mentioned sum-rules,
if we insist in obtaining results qualitatively different
from those in ref.\cite{BIM} (e.g., gauginos lighter than scalars
at the tree-level),
some scalars may get negative
mass$^2$.
This tachyonic behaviour may be just
signaling gauge symmetry breaking,
which might be a useful possibility in GUT model-building.
On the contrary, in the case of standard model 4-D strings,
the appearence of this tachyonic
behaviour could be dangerous since it could lead
to the breaking of charge and/or colour.
In order to avoid this problem, one is typically lead to a situation
with gauginos heavier than scalars, as in the overall modulus
case \cite{BIM}. We have also commented
on possible exceptions to such scenario (involving non renormalizable
Yukawa couplings or negative soft mass$^2$ for the standard
model Higgses) which could lead to scalars heavier than gauginos.
Such inversion however can take place only for special goldstino
directions, and requires necessarily a small $\sin\theta$. We recall
that the $\sin\theta \rightarrow 0$ limit was also the only one
which could produce scalars heavier than gauginos in the overall
modulus analysis, for other reasons (i.e. the different effect of
string loop corrections on gaugino and scalar masses, vanishing
at tree-level).
We have also generalized our study to include the case of orbifolds with
off-diagonal untwisted $T^{\alpha {\overline \beta}}$
moduli. In this type of models
non-diagonal metrics for the untwisted matter fields appear. In spite
of this complication, sum rules analogous to those in
eqs.(\ref{rulix},\ref{aaa}) still hold (i.e., eqs.(\ref{ddd},\ref{ccc})).
Non-diagonal metrics for the matter fields do also in general
induce off-diagonal soft-masses for the scalars which in turn can induce
flavour-changing neutral currents depending on the size of the
off-diagonal moduli, as discussed in section 3.
We have finally considered the $\mu$ and $B$ terms obtained
in orbifold schemes. We have shown that the scheme in ref.\cite{GM}
in which a $\mu $-term is generated from a bilinear piece in the
K\"ahler potential, is rather constrained in its orbifold
implementation. We find that {\it irrespective of the Goldstino
direction} one always gets $|tg\beta |=1$ at the string scale.
Another way of stating the same result is that the flat direction
$\langle H_1 \rangle =\langle H_2 \rangle $ still remains flat
after including arbitrary dilaton/moduli-induced SUSY-breaking
terms. This is an intriguing result which could have interesting
phenomenological applications. The results obtained for the
$B$-parameter in the scheme of ref.\cite{CM} in which a $\mu$-term
is generated from the superpotential are more model dependent.
A few comments before closing up are in order. First of all we are
assuming here that the seed of SUSY-breaking propagates through the
auxiliary fields of the dilaton $S$ and the moduli $T_i$ fields.
However attractive this possibility might be, it is fair to say that
there is no compelling reason why indeed no other fields in the
theory could participate. Nevertheless the present scheme
has a certain predictivity due to the relative universality
of the couplings of the dilaton and moduli. Indeed, the
dilaton has universal and model-independent couplings which are
there independently of the four-dimensional string considered.
The moduli $T_i$ fields are less universal, their number and structure
depend on the type of compactification considered. However, there are
thousands of different $(0,2)$ models with different particle content
which share the same $T_i$ moduli structure. For example, the moduli
structure of a given $Z_N$ orbifold is the same for all the thousands
of $(0,2)$ models one can construct from it by doing different
embeddings and adding discrete Wilson lines. So, in this sense,
although not really universal, there are large classes of models with
identical $T_i$ couplings.
This is not the case of generic charged matter fields
whose number and couplings are completely out of control,
each individual model being in general completely different from any other.
Thus assuming dilaton/moduli dominance in the SUSY-breaking
process has at least the advantage of leading to specific
predictions for large classes of models whereas if charged
matter fields play an important role in SUSY-breaking we
will be forced to a model by model analysis,
something which looks out of reach.
Another point to remark is that we are using the tree level forms
for both the K\"ahler potential and the gauge kinetic function.
One-loop corrections to these functions have been computed in
some classes of four-dimensional strings and could be
included in the above analysis without difficulty.
The effect of these one-loop corrections will in general
be negligible except for those corners of the Goldstino
directions in which the tree-level soft terms vanish.
However, as already mentioned above, this situation would be a sort
of fine-tuning.
More worrysome are the possible non-perturbative
string corrections to the K\"ahler and gauge kinetic functions.
We have made use in our orbifold models of the
known tree-level results for those functions.
If the non-perturbative string corrections turn out to be important,
it would be impossible to make any prediction about
soft terms unless we know all the relevant non-perturbative
string dynamics, something which looks rather remote
(although perhaps not so remote as it looked one year ago!).
One might hope that the relationships obtained among
soft terms in the dilaton/moduli dominated schemes
could be more general than the original tree-level
Lagrangians from which they are derived.
In this connection it has been recently realized that
the boundary conditions $-A=M_{1/2}={\sqrt{3}}m$
of dilaton dominance coincide with some boundary
conditions considered by Jones, Mezincescu and Yau
in 1984 \cite{JMY}. They found that those same boundary conditions
mantain the (two-loop) finiteness properties of
certain $N=1$ SUSY theories. It has
also been noticed \cite{I} that this
coincidence could be related to an underlying
$N=4$ structure of the dilaton Lagrangian
and that the dilaton-dominated boundary conditions
could also appear as a fixed point of renormalization group
equations \cite{I,J}.
This could perhaps be an indication that at least
some of the possible soft terms obtained in
the present scheme could have a more general
relevance, not necessarily linked to
a particular form of a tree level Lagrangian.
\newpage
|
\section*{\normalsize{SUMMARY}}
In this paper we outline the framework of mathematical statistics with which
one may study the properties of galaxy distance estimators. We describe,
within this framework, how one may formulate the problem of distance
estimation as a Bayesian inference problem, and highlight the crucial
question of how one incorporates prior information in this approach. We
contrast the Bayesian approach with the classical `frequentist'
treatment of parameter estimation, and illustrate -- with the simple
example of estimating the distance to a single galaxy in a redshift survey --
how one can obtain a significantly different result in the two cases. We
also examine some examples of a Bayesian treatment of distance estimation --
involving the definition of Malmquist corrections -- which have been
applied in recent literature, and discuss the validity of the assumptions
on which such treatments have been based.
\section{\normalsize{INTRODUCTION}}
Recently, the estimation of galaxy distances has assumed great importance
in cosmology. The analysis of large-scale galaxy redshift surveys, used in
conjunction with redshift-independent galaxy distance estimates, can
place powerful constraints on the values of the cosmological parameters
$H_0$ and $\Omega_0$ (c.f. Hendry, 1992b; Dekel, 1994),
and in principle can allow one to test several of the
hypotheses -- including the form of the initial spectrum of density
perturbations, the role of gravity in the growth of structure and the
clustering properties of dark matter -- on which current theories for
the formation of large scale structure in the universe are largely based.
Various methods have been
developed to reconstruct the
density and three-dimensional peculiar velocity field from galaxy
redshift and redshift-distance surveys (c.f. Dekel et al, 1990, 1993;
Simmons, Newsam \& Hendry, 1995;
Rauzy, Lachieze-Rey \& Henriksen, 1994), based upon the ansatz
that the peculiar velocity field is a {\em potential field\/} -- an idea
first developed in the {\sc{potent}} reconstruction method (Bertschinger
\& Dekel, 1989). At the same time, new statistical methods of analysing
surveys which consist of redshifts alone have been developed (c.f.
Lahav et al, 1994; Fisher et al, 1994; Heavens \& Taylor, 1995) based upon the
description of the large scale density and velocity field in terms of
sets of orthogonal functions. One of the biggest current challenges in
this field is to combine in an optimal fashion the results of these two
different methods of analysis, in order to place stronger constraints on
cosmological models and the values of cosmological parameters -- a subject
which would merit an entire article in itself.
In this article we will focus instead only on those issues which concern the
former group of reconstruction methods -- i.e. where one attempts to
obtain redshift-independent distance estimates to galaxies.
\\
Attempts to map the large scale structure of the universe from
redshift-independent galaxy distance estimates have not been without
controversy. For many years considerable debate has been generated over the
precise nature, or indeed the very existence, of galaxy concentrations such
as the `Great Attractor' in the direction of Hydra and Centaurus, for example
(Lynden-Bell et al, 1988; Dressler \& Faber, 1990;
Mathewson, Ford \& Buchhorn, 1992;
Federspiel, Sandage \& Tammann, 1994). A
significant factor fuelling this controversy has been disagreement not
so much over the astrophysical problems of determining `good' galaxy
distance indicators
(although this has undoubtedly played a part also) but rather disagreement
over the equally fundamental question of what {\em statistical\/} methods
one should adopt to analyse the galaxy data. In this paper we attempt to
clarify and place in the open some of the different statistical approaches
which have been adopted in this field of cosmological research, and to
discuss -- within the framework of mathematical statistics -- the different
underlying philosophies upon which (often implicitly) they are based. Our
discussion should be viewed as a general introduction to the problem,
suitable for a reader previously unfamiliar both with the relevant
astronomical details of measuring galaxy distances and with the basic
theory of probability and statistics upon which the topic is founded.
References to more detailed articles, covering both the astronomical and
statistical aspects of the problem, will be given wherever appropriate.
\\
The measurement of the distance of a galaxy, is an example of an
{\em inference\/} problem: i.e. one cannot measure the distance directly
but must infer it from the measurement of some other physical
characteristic, such as the apparent visual magnitude or angular diameter.
If one knew precisely the {\em absolute\/} magnitude or intrinsic
diameter of the galaxy then one could immediately arrive at an exact
determination of the galaxy distance. In early studies of the large scale
distribution and motion of galaxies (c.f. Rubin et al, 1976; Sandage \&
Tammann, 1975a,b) the approach was simply to assume {\em a priori\/} some
fiducial value for this absolute magnitude or diameter and thus infer
galaxy distances on that basis. In practice, however, not all galaxies have
the same absolute magnitude or diameter and so the inference is
statistical in nature. Shortly after these early studies
significant progress was made with the identification of empirical
relationships between absolute magnitude and diameter and other,
distance-independent but directly measurable, physical quantities such as
velocity dispersion or colour (c.f. Faber \& Jackson, 1976; Tully \&
Fisher, 1977; Visvanathan \& Sandage, 1977).
The Tully-Fisher relation, for example, essentially expresses a power law
relationship between the luminosity and the rotation velocity -- as measured
from e.g. the 21cm neutral hydrogen radio emission -- of spiral galaxies.
Thus one measures the 21cm line width of neutral hydrogen for a given
spiral galaxy, applies the Tully-Fisher relation to infer the
absolute magnitude of the galaxy, and then infers the galaxy distance from
its observed apparent magnitude.
\\
In the past decade the Tully-Fisher, and other similar,
relations have been
further refined and placed upon a firmer theoretical footing,
(Pierce \& Tully, 1988; Salucci, Frenk \& Persic, 1993; Hendry et al, 1995)
but they
still contain a significant degree of intrinsic scatter and so do not
provide an exact determination of absolute magnitude or diameter.
Hence, the galaxy distance inferred from such a relation is still
inherently statistical. In the language of mathematical statistics, the
intrinsic scatter of the relation means that we can construct only an
{\em estimator\/} of the galaxy distance, and that distance estimator will
itself be subject to error. More formally, the distance estimator is a
random variable with a definite distribution function, or equivalently
probability density function (pdf), and {\em a fortiori\/} mean and
variance.
\\
Unfortunately there is no unique way to construct distance estimators.
One can make a choice of distance estimator which has certain desirable
properties, the most obvious being that its distribution should have
a small `spread', or variance; on average over many realisations the
estimator should give the true distance of the galaxy; and the
estimator should use all of the information about the galaxy
distance available in the data. These rather loosely defined properties
have their corresponding rigorous definitions in the statistical
literature, and these are referred to as {\em efficiency\/},
{\em unbiasedness\/} and {\em sufficiency\/} respectively.
\\
One should remark that when measurement errors and
intrinsic variability are small in the physical system which one is
modelling, then the adoption of a broad class of different statistical
methods -- or even different statistical philosophies -- in testing models
from observational data will usually make little difference to one's
conclusions. Large discrepancies in the conclusions reached by various
authors in the literature concerning the estimated distances of galaxies
and clusters therefore arise primarily because of large intrinsic
uncertainties inherent to the data.
In other words, galaxy distance indicators are {\em noisy\/}, with typical
distance errors from, e.g., the Tully-Fisher relation of around $20\%$ or
larger to individual galaxies. It is this fact which makes the question of
how one approaches the problem of choosing the `best' galaxy distance
estimator a non-trivial, and an extremely important, one. The typical size
of distance errors has led many cosmologists to attempt to incorporate
prior information on the distribution of galaxy distances when defining
distance estimators, with the aim of reducing the uncertainty in the
final estimate. All examples of this approach can be traced back to what
is termed in the statistical literature as a {\em Bayesian\/} treatment of
the problem of distance estimation, although references in the cosmology and
astronomy literature have often not explicitly used the term `Bayesian',
nor indeed used wholly orthodox Bayesian methods, in their description
of the problem. There are indeed some difficulties with this approach.
One the one hand there are philosophical and methodological problems
that have long been recognised and debated by statisticians
(c.f. Kendall \& Stuart, 1963; Mood \& Graybill, 1974; von Mises, 1957;
Feigelson \& Babul, 1992) which
go to the root definitions and concepts in the theory of probability.
On the other, there is often no clear-cut way of deciding upon the
nature of prior information one can justifiably use. This paper is not
the appropriate place to discuss either of these questions in any great
depth. We would like to emphasise here, however, the principle employed
in Bayesian inference problems in the general statistics literature:
that results which depend heavily on the choice of prior information
should be treated with caution.
\\
Whilst the problem of galaxy distance estimation raises certain
statistical issues which are somewhat unique to astronomy --
in particular the important role of observational selection effects and the
modelling of the physical processes underlying the various distance
relations which are applied to galaxies -- the fundamental concepts are
{\em precisely\/} the same as one finds in the general statistical
literature on inference problems and estimation. It seems sensible,
therefore, for cosmologists to make full use of the `machinery' --
the definitions, notation and general results -- developed by
statisticians for tackling such problems. In this paper, as in our
earlier papers on this subject, we shall attempt to adhere to this
practice.
\\
The structure of this paper is as follows. In section 2 we discuss
in more detail the nature of distance estimators, placing our
discussion in the rigorous context of mathematical statistics and
introducing the appropriate notation and conventions. We go on to
discuss the role of prior information, to explain the concepts of
a `Bayesian' approach to estimation problems, and to examine the
relationship between Bayesian and more orthodox or `frequentist'
approaches. We show, by means of the simple example of estimating
the distances to galaxies in a single catalogue, how a Bayesian
and frequentist approach will yield different results. In section 3
we discuss the various galaxy distance estimators which have been
used in recent literature, drawing particular attention to the
statistical `philosophy'
(i.e Bayesian or frequentist) upon which
they are based, the validity of the assumptions inherent in their
definition, and the extent to which they can be regarded as `good'
estimators -- in the sense of e.g. unbiasedness, efficiency and
sufficency, as introduced above. Finally we discuss the practical
outcomes of using these different estimators for determining
distances to individual galaxies and clusters and in the analysis of
the peculiar velocity and density field by, e.g., the {\sc{potent}}
based methods mentioned above.
\section{\normalsize{STATISTICAL PROPERTIES OF DISTANCE ESTIMATORS}}
One of the purposes of this section is to clarify our notation and
statistical approach for the benefit of the reader previously
unacquainted with the general statistics literature. In the interests
of brevity we shall present here only the essential ideas and omit
unnecessary detail, perhaps at the risk of appearing simplistic. A
more thorough, and wholly rigorous, treatment of the mathematical
foundations of parameter estimation can be found in a large number
of textbooks on probability and statistics (c.f. Hoel, 1962;
Kendall \& Stuart, 1963; Mood \& Graybill, 1974; Hogg \& Craig, 1978)
\section*{\normalsize{What Is an Estimator?}}
In rough terms, an estimator of some unknown parameter is a rule based
on statistical data -- i.e. a random sample drawn from some underlying
population -- for estimating the value of that parameter. If the
parameter of interest is $q$ then we shall write ${\hat{\bf{q}}}$ to
denote an estimator of $q$, following the standard
statistical convention. Note that ${\hat{\bf{q}}}$ is written
in bold face to indicate the fact that it is a {\em random\/}
or {\em statistical variable\/}
(since it is a function of data which are themselves statistical
variables), again in keeping with the standard practice in the
literature. One cannot, of course, expect ${\hat{\bf{q}}}$ to take
on the true value of $q$, $q_0$ say, for {\em every\/} set of statistical
data, but we would regard an estimator as `good' if it tends to yield the
value $q_0$ `on average', or `in the long run' -- rather vague
statements which can be quantified in terms of the {\em bias\/} and
{\em loss function\/} associated with the estimator chosen, as we
discuss below.
\\
By way of an illustrative example, a simple galaxy distance
estimator could be constructed only from the
obaserved apparent magnitude of a galaxy (c.f. Hendry \& Simmons, 1990;
Hendry, 1992a). Thus we may write
\begin{equation}
{\bf{m}} - {\bf{M}} = 5 \log r + 25
\end{equation}
where $r$ is the true distance, measured in Mpc, and ${\bf{m}}$
and ${\bf{M}}$ denote
the apparent and absolute magnitude of the galaxy respectively. Of course
the actual distance of the galaxy can only be obtained if there is no
error on the measured value of ${\bf{m}}$ and if ${\bf{M}}$ is known. We can
estimate $r$, however, by making some assumption about the value of
${\bf{M}}$ (for
simplicity we shall ignore any error on ${\bf{m}}$ in this discussion) and
solving for $r$ in equation (1). Suppose we take the value of ${\bf{M}}$ to
be the mean value of absolute magnitude, $M_0$ say, for the underlying
population of all galaxies of a certain Hubble type. We thus obtain an
estimator of log distance, viz
\begin{equation}
{\widehat{\log {\bf{r}}}} = 0.2 ( {\bf{m}} - M_0 - 25)
\end{equation}
Here the hat indicates an estimator. If we consider that the
galaxy has been randomly selected from an imaginary population of
galaxies all at the same distance, but with different absolute
magnitudes, then ${\widehat{\log {\bf{r}}}}$ must be considered to be
a statistical variable, as noted previously. The statistical properties
of ${\widehat{\log {\bf{r}}}}$ depend on the galaxy luminosity function
and on the selection function which determines whether a galaxy will or
will not be observed at true distance, $r$. It follows from equations
(1) and (2) that we may write
\begin{equation}
{\widehat{\log {\bf{r}}}} = \log r + 0.2 ( {\bf{M}} - M_0 )
\end{equation}
For brevity we shall in future refer to $\log r$ as $w$,
and ${\widehat{\log {\bf{r}}}}$ as ${\hat{\bf{w}}}$.
\\
In general, the underlying pdf {\em before\/} selection for
${\bf{M}}$ is not known. This pdf is usually assumed to be independent
of position and is just the luminosity function (LF) of ${\bf{M}}$,
written $\Psi({\bf{M}})$. The distribution,
$\Psi_{\rm{obs}}({\bf{M}}|r)$,
of ${\bf{M}}$ for observable galaxies at actual distance, $r$,
will depend upon the
selection functio\rm{n a}nd indeed also on $r$ (although for simplicity
we assume here no dependence on direction). Once this pdf,
$\Psi_{\rm{obs}}({\bf{M}}|r)$, is given the pdf of any function of the
random variable, ${\bf{M}}$, may be determined. In particular the pdf
of ${\hat{\bf{w}}}$ defined by equation (3) may easily be found.
Note that while ${\hat{\bf{w}}}$ itself does not depend on
$w$, the pdf of ${\hat{\bf{w}}}$ {\em does\/} depend upon
the true value of the parameter, as one might expect.
\section*{\normalsize{Biased and Unbiased Estimators}}
The mean, or {\em expected\/}, value of a random variable associated with
a galaxy may be taken with respect to either the observable
or the intrinsic galaxy distribution. We shall almost invariably
consider the expectation with respect to the observable distribution in
this paper. Thus the estimator of log distance, ${\hat{\bf{w}}}$,
is defined to be unbiased if
\begin{equation}
E ( {\hat{\bf{w}}} | w ) = w
\end{equation}
where the expectation value of any function, $f({\bf{M}})$, of
${\bf{M}}$ is defined as
\begin{equation}
E [ f({\bf{M}}) | w ] = \int \! f({\bf{M}})
\Psi ({\bf{M}} | w) d{\bf{M}}
\end{equation}
The bias, $B(w)$, is defined as
\begin{equation}
B(w) = E [ {\hat{\bf{w}}} | w ] - w
\end{equation}
When a galaxy survey is subject to a selection limit on apparent
magnitude, the estimator of log distance given by equation (2)
is {\em biased\/} for all true log distances. Moreover, simply
replacing the mean absolute magnitude, $M_0$, of the underlying
population by some fiducially corrected value, $M_0 + c$, where
$c$ is a constant, cannot eradicate this bias (c.f. Hendry \&
Simmons, 1990). One can apply an iterative procedure --
effectively adding a non-constant correction to $M_0$ -- which
considerably reduces the bias of ${\hat{\bf{w}}}$, although
this procedure does not converge to an unbiased estimator for
all log distance (Hendry, 1992a). It has been shown,
however, (c.f. Schechter, 1980; Hendry \& Simmons, 1994) that
in the case of a relation of Tully-Fisher type -- where one
has an additional observable correlated with absolute magnitude --
if the second observable is free from selection effects then one
{\em can\/} define an estimator which is unbiased at all true
log distances. We return to this issue in section 3.
\section*{\normalsize{Minimum Variance and Efficient Estimators}}
There are obvious advantages in using unbiased estimators: in particular,
for large samples -- e.g. when one is estimating the distance to a
rich cluster of galaxies -- the mean estimated distance for the sample
will also be unbiased, and of course will have decreasing variance as the
sample size increases.
Furthermore, if we are interested in, say, the distribution
of actual distances of a catalogue of galaxies, the
histogram of {\em estimated\/} distances can be readily
deconvolved to yield an estimate of this underlying
distribution of true distances. For biased
estimators this would be more difficult (c.f. Eddington, 1913; Newsam,
Simmons \& Hendry, 1994, 1995). Similarly, in model fitting problems -- the
simplest of which in the present context is e.g. the determination of
the Hubble constant -- we can expect parameter estimation to be much easier
if we begin with unbiased estimators. Unbiasedness is not the only criterion
for choosing an estimator, however. It is also natural to desire the
estimator to have a small variance. The variance, $V(w)$, of an estimator is
defined as
\begin{equation}
V(w) = E [ ( {\hat{\bf{w}}} - w )^2 | w ]
\end{equation}
In practice one finds that there is a trade-off between small variance and
small bias, in the sense that if you reduce one then you increase the other.
The Cramer-Rao inequality places a lower bound on the variance for both
biased and unbiased estimators (c.f. Hogg and Craig, 1978; Hendry, 1992a;
Gould, 1995; Zaccheo et al, 1995), and an efficient
estimator is one which {\em attains\/} that lower bound -- i.e. which
is a minimum variance estimator.
\\
In choosing an estimator it is also usually convenient to
introduce a {\em loss function\/}, which essentially quantifies
the `loss', or cost, of making an incorrect estimate of a
parameter. An obvious loss function to consider is
\begin{equation}
L({\hat{\bf{w}}},w) = ({\hat{\bf{w}}} - w)^2
\end{equation}
A good estimator should yield low values of the expected loss
for a large range of values of the parameter $w$. This
expected loss is called the {\em risk\/}, i.e.
\begin{equation}
R(w) = E [ L({\hat{\bf{w}}},w) ]
\end{equation}
Note that for an unbiased estimator the risk and variance are identical,
but for a biased estimator the risk is always strictly greater than the
variance. Thus, if one has an estimator with small variance but large bias,
this would still result in an estimator of large risk -- indicating that
risk is often the more meaningful quantity in comparing estimators. In
general the bias, variance and risk of an estimator are related by the
following simple expression
\begin{equation}
R(w) = V(w) + [ B(w)]^2
\end{equation}
\section*{\normalsize{Sufficiency}}
In estimating the distance of a galaxy one does not generally adopt an
estimator of the simple form of equation (2), which is a function only of
apparent magnitude, but rather makes use of a distance indicator such as the
Tully-Fisher relation which depends upon the the strong correlation between
absolute magnitude and some other distance-independent, directly measurable,
observable. Since the underlying physical relationship in an indicator of this
type is unlikely to depend upon only two variables, one could in principle
construct a distance estimator as a function of an arbitrary number of
observables, or {\em statistics\/}. The bias and risk of such an estimator
would depend, of course, upon how well correlated were the observables. In
Hendry \& Simmons (1994) the general case of estimators formed from three
correlated observables is formulated, and in Kanbur \& Hendry (1995) a
specific example is considered where the addition of a {\em fourth\/}
observable -- the maximum apparent magnitude -- to the period, mean luminosity,
colour relation for Cepheid variable stars does indeed result in a distance
estimator of significantly smaller variance and risk.
\\
A obvious general question to ask, then, is whether there exists a function,
say ${\hat{\bf{w}}}({\bf{x}}_1,...,{\bf{x}}_n)$, of a set of observables,
${\bf{x}}_1,...,{\bf{x}}_n$, which `contains' all of the information about
the true value of $w$. Such a function is known as a {\em sufficient
statistic\/} -- and hence would define a sufficient estimator -- for $w$, and
so should be preferred over another estimator without this property. The
property of sufficiency
can be given a more rigorous mathematical definition in terms of the joint
pdf of ${\bf{x}}_1,...,{\bf{x}}_n$ and ${\hat{\bf{w}}}$ (c.f. Mood \&
Graybill, 1974). Suppose that ${\hat{\bf{w}}}_*({\bf{x}}_1,...,{\bf{x}}_n)$
is another statistic based on the observables,
${\bf{x}}_1,...,{\bf{x}}_n$, which is not a function of
${\hat{\bf{w}}}$. Then ${\hat{\bf{w}}}$ is defined to be sufficient if, for
{\em any\/} such ${\hat{\bf{w}}}_*$, the conditional distribution of
${\hat{\bf{w}}}_*$ given ${\hat{\bf{w}}}$ does {\em not\/} depend on the
true parameter value, $w$.
\\
This definition essentially states that once the value of the sufficient
statistic has been specified, one cannot find any other statistic based on
the same set of observables which gives any further information about the
true value of $w$. In a sense, ${\hat{\bf{w}}}$ `exhausts' all the
information about $w$ that is contained in the observed values of
${\bf{x}}_1,...,{\bf{x}}_n$.
\section*{\normalsize{Bayes' Estimators}}
So far we have said nothing about the incorporation of
prior information in the estimation of galaxy distances. Bayesian
approaches attempt to do precisely this.
\\
A fully fledged Bayesian approach would regard $w$ -- in the above
notation -- not as a parameter, but as a statistical variable. The
probability (more commonly referred to as the {\em likelihood\/}) of this
variable taking any given value would be determined by what is
known as its {\em prior distribution\/}: prior, that is, to the data
that we presently have at hand. In the cosmological setting, therefore,
${\bf{w}}$ -- the log distance of a galaxy -- would be taken to have a prior
distribution before the apparent magnitude or diameter or line width
of this galaxy were measured. This prior distribution would be
based on previous information about the distribution of
galaxies as a whole -- or even preconceptions about this
distribution, such as the assumption that the spatial distribution of
galaxies be uniform. In this case one has to modify the orthodox frequentist
view of probability as a `limit' of relative frequencies and adopt instead a
view of probability as a measure of one's state of knowledge about a random
variable.
\\
The {\em posterior\/} distribution for ${\bf{w}}$, once the data for
a particular galaxy has been taken into account, is then
obtained by applying Bayes' theorem. Suppose one's distance estimator is a
function of two variables, ${\bf{m}}$ and ${\bf{P}}$ -- denoting for example
apparent magnitude and log rotation velocity for the Tully Fisher
relation. Bayes' theorem states that
\begin{equation}
p({\bf{m}},{\bf{P}}|{\bf{w}} ) p({\bf{w}}) =
p({\bf{w}}|{\bf{m}},{\bf{P}})p({\bf{m}},{\bf{P}})
\end{equation}
Taking $p({\bf{m}},{\bf{P}})$ to pe a constant, one obtains the
posterior distribution for ${\bf{w}}$, viz
\begin{equation}
p({\bf{w}}|{\bf{m}},{\bf{P}}) = C \, p({\bf{m}},{\bf{P}}|{\bf{w}})
p({\bf{w}})
\end{equation}
where $p({\bf{w}})$ is the prior, $C$ is a normalisation constant and
$p({\bf{m}},{\bf{P}}|{\bf{w}})$ is the conditional probability of
${\bf{m}},{\bf{P}}$ given ${\bf{w}}$.
\\
This approach in itself does not give an estimator of
${\bf{w}}$, which is a statistical variable and
not strictly speaking a parameter in the Bayesian
context, but rather it gives a posterior pdf for ${\bf{w}}$ from which one
may define a {\em Bayes' estimator\/} (c.f. Mood \& Graybill, 1974)
in the following way. A Bayes' estimator, ${\bf{\hat{w}}}_{\rm{bayes}}$
minimises the risk, $R({\bf{w}})$ averaged over the prior
distribution, $p({\bf{w}})$ for ${\bf{w}}$. Thus for a
Bayes estimator the integral
\begin{equation}
\int \! R({\bf{w}}) \, p({\bf{w}}) \, d{\bf{w}}
\end{equation}
is a minimum. It can be shown that a Bayes' estimator in fact minimises
the loss function averaged over the distribution for ${\bf{w}}$
conditional on the observed data. Explicitly it minimises
\begin{equation}
\int \! L({\bf{\hat{w}}}_{\rm{bayes}},{\bf{w}})
p({\bf{w}}|{\rm{data}})d{\bf{w}}
\end{equation}
from which ${\bf{\hat{w}}}_{\rm{bayes}}$ can be found.
\\
It is instructive to consider a simple example where we are
estimating the log distance, $w$, of a galaxy. Let
us assume that we have already an unbiased (in the sense of equation 6)
`raw' estimator, ${\bf{\hat{w}}}$, based on some distance
indicator, which we shall for expediency take to be
normally distributed about the true log distance ${\bf{w}}$ with variance
$\sigma^2$.
Thus the conditional distribution for ${\bf{\hat{w}}}$ given ${\bf{w}}$ is
\begin{equation}
p({\bf{\hat{w}}}|{\bf{w}}) = \frac{1}{\sqrt{2 \pi \sigma}}
\exp [ - \frac{1}{2} (\frac{{\bf{\hat{w}}} - {\bf{w}}}{\sigma})^2 ]
\end{equation}
Let us assume, however, that the galaxy is randomly
selected from some underlying population with true log
distance ${\bf{w}}$ distributed normally about some mean value,
$w_c$, and variance $\sigma^2_c$ -- where the
subscript $c$ refers to the catalogue from which the
galaxy is drawn. This normal distribution is taken to be the prior, so in
the above notation
\begin{equation}
p({\bf{w}}) = \frac{1}{\sqrt{2 \pi \sigma_c}}
\exp [ - \frac{1}{2} (\frac{{\bf{w}} - w_c}{\sigma_c})^2 ]
\end{equation}
It is now straightforward to show that the conditional distribution
for ${\bf{w}}$ given the value of ${\bf{\hat{w}}}$ is
normally distributed with mean, ${\bf{w}}_B$ and variance
$\sigma^2_B$ given by
\begin{equation}
{\bf{w}}_B = {{{\bf{\hat{w}}} + \beta w_c} \over {1+\beta}}
\end{equation}
and
\begin{equation}
{\sigma^2_B} = {{\sigma^2} \over {1+\beta}}
\end{equation}
where $\beta = \sigma^2 / \sigma^2_c$,
from which it follows that
\begin{equation}
{\bf{\hat{w}}}_{\rm{bayes}} = {{{\bf{\hat{w}}} + \beta w_c} \over
{1+\beta}}
\end{equation}
The interpretation of this result is very
straightforward. If the variance of the indicator is much
smaller than the population variance of the normal distribution of
true log distance for observable galaxies then $\beta \simeq 0$ and
one obtains essentially the `raw' log distance estimator,
${\bf{\hat{w}}}$, suggested by the indicator. If, on the other
hand, the indicator provides very poor information about
the distance of the galaxy then $\beta$ is very large,
and the Bayes estimator yields approximately $w_c$, the mean true log
distance of the observable galaxies in the catalogue. This simple example
demonstrates that, provided the scatter in one's distance indicator is
sufficiently small, one obtains essentially the same estimator irrespective
of whether one adopts a Bayesian approach or not -- and the estimator is
thus largely insensitive to the prior information. The role of the prior
becomes increasingly important, however -- and the difference between a
Bayesian and frequentist approach becomes more apparent -- as the scatter
in the distance indicator increases.
\\
One can regard equation (19) as defining a {\em correction\/}
to ${\bf{\hat{w}}}$ based on the prior information -- in this case that the
underlying populatoin of true log distance is normally distributed.
Corrections of this type
have come to be known in the cosmology literature as {\em Malmquist\/}
corrections, and in the context of mapping large scale structure they were
initially applied assuming the distribution of galaxies to be spatially
homogeneous (c.f. Lynden-Bell et al, 1988; Dekel et al, 1993) -- just as
the distribution of {\em stars\/} had been assumed homogeneous in the
original analytical treatments of Malmquist (1920, 1922) and
Eddington (1913). Recently, however, attempts have been made to apply more
general, inhomogeneous, Malmquist corrections which address the fact that
the galaxy distribution displays small-scale clustering
(c.f. Landy \& Szalay, 1992; Hudson, 1994; Dekel, 1994; Newsam, Simmons \&
Hendry, 1995; Hudson et al, 1995; Freudling et al, 1995).
We briefly consider some important technical problems
regarding the application of inhomogeneous Malmquist corrections in
section 3. It is worth noting here, however, that an entirely frequentist
approach to distance estimation has the advantage that the definition
of an unbiased estimator is completely independent of the underlying
galaxy true number density -- and hence is unaffected by arguments about
the form of prior distribution which one should adopt.
\section{\normalsize{GALAXY DISTANCE INDICATORS IN RECENT LITERATURE}}
Most redshift-independent methods of estimating galaxy distances which have
featured in the recent cosmological literature are based upon
{\em secondary\/} distance indicators -- which require to be calibrated using
a sample of galaxies in, e.g., a nearby cluster, the distance of which is
already known. Notable exceptions to this have been the recent extension
to beyond the Local Group of the extragalactic distance scale measured from
Cepheid variables and the application of the expanding photosphere method
(EPM) to determine the distances of type II supernovae (SN). Both Cepheids
and type II SN are examples of primary distance indicators which can be
calibrated either locally -- within our own galaxy -- or from theoretical
considerations. For a discussion of the physical basis for these
indicators the reader is referred to, e.g., Kirschner \& Kwan (1974),
Eastman \& Kirschner (1989), Jacoby et al (1992) and references therein. Both
indicators have a small intrinsic dispersion ($\sim 10 - 15\%$ to individual
objects) and are thus considerably less susceptible to the problems which
arise in the definition of Malmquist corrections and sensitivity to the
choice of
prior information (essentially because the $\beta$, the ratio of the estimator
variance to the variance of the underlying population, is small).
This property of course makes both Cepheids and type II SN well suited to the
estimation of the Hubble constant -- either directly or in combination with
other secondary indicators such as type Ia SN (c.f. Saha et al, 1994; 1995).
Indeed, the high estimates of H$_0$ reported in
Freedman et al (1994) and Pierce et al (1994), based on the distance of
Cepheids in Virgo cluster galaxies, and those of Schmidt et al (1992, 1994)
based on the EPM distances of type II SN beyond the Local Supercluster
(and thus less adversely affected by peculiar velocities), provide a
compelling argument in favour of a value of H$_0 \geq
60$kms$^{-1}$Mpc$^{-1}$ -- despite the difficulties of reconciling
these results with astrophysical estimates of the age of the galactic disc
(c.f. van den Bergh 1995; Chamcham \& Hendry, 1995).
\\
Of the secondary distance indicators currently in widespread use, only two
are thought to be sufficiently accurate to make the question of how to best
use prior information essentially unimportant: these are surface brightness
fluctuations (SBF) and the luminosity -- light curve shape relation for
type Ia SN. The former distance indicator, SBF, was pioneered by
Tonry \& Schneider (1988) and is based upon the fact that the
fluctuations -- due to the discreteness of individual
stars -- in surface brightness across the CCD image of a nearby elliptical
galaxy will be larger than those for a more distant galaxy. The physical
basis of SBF and the details of its calibration are described in
Jacoby et al (1992). Relative distances of a typical accuracy of
5\% have been derived to a sample of several hundred ellipticals out to a
redshift of around 6000 kms$^{-1}$ using this indicator (Dressler, 1994).
\\
Type Ia SN have long been recognised as useful `standard candles' since they
are observable to very large distances and have a luminosity function which is
well described a Gaussian distribution of dispersion around 0.5 mag.
(Sandage \& Tammann, 1993; Hamuy et al, 1995). In Vaughan et al (1995), it is
argued that the pre-selection of SN based on a colour criterion reduces this
dispersion to $\sim 0.3$ mag., which -- although a significant improvement --
still represents a typical percentange distance error of around 15\% to an
individual galaxy. In Riess, Press \& Kirschner (1995a,b) however, the
shape of the SN light curve is used to more tightly constrain the peak
luminosity and leads to a typical relative distance error of only 5\% -- small
enough to render Malmqust corrections largely unimportant. This method has been
used both to estimate the Hubble constant and to determine the bulk flow
motion of the Local Group on a scale of $\sim 7000$ kms$^{-1}$, yielding a
motion which is consistent with the COBE measurement of the dipole
anisotropy in the microwave background radiation, but inconsistent with the
dipole motion reported by Lauer \& Postman (1994), based on the redshifts of
Abell clusters at distance of $8000 - 11000$ kms$^{-1}$.
\\
The vast majority of recent analyses of the peculiar velocity and density
fields, and the estimation of the density parameter $\Omega_0$ using
redshift-independent distance indicators, have
been carried out primarily with the Tully-Fisher (TF) and $D_{\rm{n}}-\sigma$
distance indicator relations for spirals and ellipticals respectively. As we
remarked above, the TF relation essentially expresses a power
law relationship between the luminosity and rotation velocity for spiral
galaxies; the $D_{\rm{n}}-\sigma$ relation similarly expresses a power law
relationship between the central velocity dispersion and isophotal
diameter of early-type galaxies (c.f. Jacoby et al, 1992). Although the
number of galaxy distances estimated by these two relations currently
stands at over 4000 (around a factor of ten larger than the number of
distance estimates from SBF and SN distance indicators), and continues
to grow rapidly each year, both the TF and $D_{\rm{n}}-\sigma$ relations are
considerably more noisy -- with dispersions of around 20\% to individual
galaxies. It is for this reason that the issue of how -- or
indeed if -- one should make use of prior information in the definition of
`optimal' estimators continues to be regarded as of crucial importance when
interpreting the results of applying these distance indicators to analyse
redshift surveys.
\\
Both the TF and $D_{\rm{n}}-\sigma$ relations are usually calibrated by
performing a linear regression on a calibrating sample of
galaxies whose distances are otherwise known. It is instructive to
consider this calibration procedure in more detail, in order
to illustrate some of the statistical pitfalls which may arise, for the
generic example of the TF relation. As before, we denote the log
rotation velocity by ${\bf P}$ and let ${\bf \hat{M}}$ denote the
estimator of absolute magnitude which one derives from the TF relation, from
which one may derive the corresponding `raw' estimator of log distance,
${\hat{\bf{w}}}$, from equation (3) in the obvious way.
Thus, we obtain from the calibration a linear relationship between
${\bf \hat{M}}$ and ${\bf P}$,
\begin{equation}
\label{eq:Mhat}
{\bf \hat{M}} = \alpha {\bf P} + \beta
\end{equation}
where $\alpha$ and $\beta$ are constants. The choice of {\em which\/} linear
regression is most appropriate is non-trivial when one's survey is subject to
observational selection effects. We can demonstrate this with the following
simple example. Suppose that the intrinsic joint distribution of absolute
magnitude and log(rotation velocity) is a bivariate normal.
Figure 1 shows schematically the scatter in the TF relation in
this case, for a calibrating sample which is free from selection effects --
e.g. a nearby cluster. (More precisely, the ellipse shown is an
isoprobability contour enclosing a given confidence region for ${\bf M}$ and
${\bf P}$). The solid and dotted lines show the linear relationship obtained by
regressing rotation velocities on magnitudes and magnitudes on line widths
respectively.
Thus the dotted line is defined as the expected value of ${\bf M}$ at given
${\bf P}$, while the solid line is defined as the expected value of ${\bf P}$
at given ${\bf M}$. Since in practice one wishes to infer the value of
${\bf M}$ from the measured value of ${\bf P}$, the ${\bf M}$ on ${\bf P}$
regression has been referred to in the literature as defining the `direct'
or `forward' TF relation, while using the ${\bf P}$ on ${\bf M}$ regression
defines the `inverse' TF relation. For the bivariate normal case the
equations of the direct and inverse regression lines are as follows:-
\begin{equation}
\label{eq:MgivP}
E ({\bf M}|{\bf P}) = \rmsub{M}{0} + \frac{\rho \rmsub{\sigma}{M}}
{\rmsub{\sigma}{P}} ({\bf P} - \rmsub{P}{0})
\end{equation}
\begin{equation}
\label{eq:PgivM}
E ({\bf P}|{\bf M}) = \rmsub{P}{0} + \frac{\rho \rmsub{\sigma}{P}}
{\rmsub{\sigma}{M}} ({\bf M} - \rmsub{M}{0})
\end{equation}
where $\rmsub{M}{0}$, $\rmsub{P}{0}$, $\rmsub{\sigma}{M}$, $\rmsub{\sigma}{P}$
and $\rho$ denote the means, dispersions and correlation coefficient of the
bivariate normal distribution of ${\bf M}$ and ${\bf P}$.
Both regression lines can be written in the form of equation (20),
thus defining ${\bf \hat{M}}$ as a function of ${\bf P}$, although of
course the constants $\alpha$ and $\beta$ are different in each case.
Moreover the definition of ${\bf \hat{M}}$ is subtly different in each case.
For the direct regression ${\bf \hat{M}}$ is identified as the mean
absolute magnitude at the observed log line width. For the inverse regression
on the other hand ${\bf \hat{M}}$ is defined such that the observed log line
width is equal to its expected value when ${\bf M} = {\bf \hat{M}}$.
Consequently, as is apparent from their slopes, the direct and inverse
regression lines give rise to markedly different distance
estimators, although it is straightforward to show that in the absence of
selection effects both estimators are unbiased, in the sense defined in
equation (4), above.
\\
\begin{figure}
\vspace{7cm}
\caption{Schematic `Direct' and `Inverse' Tully-Fisher relations for the
case of a nearby, completely sampled, cluster.}
\end{figure}
The situation is very different when we include the effects of observational
selection, however. This is illustrated in Figure 2, which
shows the scatter in the TF relation in a calibrating sample subject to a
sharp cut-off in absolute magnitude -- as would be the case in e.g. a
distant cluster observed in an apparent magnitude-limited survey. We can
see that in this case the slope of the direct regression
of ${\bf M}$ on ${\bf P}$ is substantially changed from that in the nearby
cluster -- indeed the direct regression is no longer linear at all. This means
that if one calibrates the TF relation in the nearby cluster using the direct
regression and then applies this relation to the more distant cluster, one
will systematically underestimate its distance, since the expected value of
${\bf M}$ given ${\bf P}$ in the distant cluster is systematically brighter
than that in the nearby cluster as fainter galaxies progressively `fade out'
due to the magnitude limit. The corresponding `direct', or `M on P', log
distance estimator will therefore be negatively biased.
\\
\begin{figure}
\vspace{7cm}
\caption{Schematic `Direct' and `Inverse' Tully-Fisher relations for
the case of a distant cluster subject to a sharp selection limit on
absolute magnitude.}
\end{figure}
In an important paper Schechter (1980) observed that the slope of the
inverse regression line is unchanged,
irrespective of the completeness
of one's sample, provided that the selection effects are in magnitude only.
We can see that this observation is valid in the simple case considered in
Figure 2. In other words the expected value of ${\bf P}$ given ${\bf M}$ is
unaffected by the selection effects and, therefore, defines an unbiased log
distance estimator. In Hendry \& Simmons (1994), Schechter's result is derived
within the rigorous framework of mathematical statistics, and the
assumptions upon which it is based are generalised. In particular it is
shown that the inverse TF log distance estimator is gaussian and unbiased at
all true log distances provided only that the conditional distribution of
${\bf P}$ given ${\bf M}$ is Gaussian, that $E({\bf P}|{\bf M})$ is a linear
function of ${\bf M}$, and that the sample is not subject to selection on
rotation velocity. Moreover, since the inverse log distance estimator is
Gaussian it will also automatically be a {\em sufficient\/} and
{\em efficient\/} estimator, as defined in section 2. In Hendry (1992a)
it was also shown that when there is no selection on rotation velocity
then the inverse log distance estimator is the {\em only\/} unbiased
estimator which is a linear function of log rotation velocity and apparent
magnitude. In particular, the `orthogonal'
(c.f. Giraud 1987), `bisector' (c.f. Pierce \& Tully 1988) and
`mean' (c.f. Mould et al 1993) regression lines also give rise to estimators
which are biased at all true log distances in this case. A similar
conclusion was also reached in Triay, Lachieze-Rey \& Rauzy (1994).
\\
The unbiased properties of the inverse TF relation have led to its use
in defining a `raw' distance estimator in
a number of different recent analyses of the peculiar velocity field,
including Newsam, Simmons \& Hendry (1995), Freudling et al (1995),
Nusser \& Dekel (1994), Shaya, Tully \& Pierce (1992) and Shaya, Tully \&
Peebles (1995). Its acceptance has been far from universal, however. Part
of the reason for this is that, of course, in practice it is {\em not\/}
the case that galaxy samples are free from selection effects on rotation
velocity. In fact, it is commonly the case that redshift surveys are first
selected on the basis of either apparent diameter or B-band apparent
magnitude, or both, while the TF photometry is then carried out in the
near infra-red, or I-band. This leads to a considerably more complex
selection function, as modelled in Sodre \& Lahav (1993), which in general
renders {\em all\/} linear regressions biased. Essentially this problem
arises because diameter, I-band and B-band magnitude and rotation velocity
are mutually correlated variables, so that the selection on B-band
magnitude and angular diameter `pollutes' the joint distribution of
I-band magnitude and rotation velocity in the TF relation -- thus
effectively rendering the assumptions inherent in deriving the unbiasedness
of the inverse TF relation no longer valid (c.f. Hendry \& Simmons 1994;
Willick 1994).
\\
One {\em can\/} determine the correct slope and zero point of the `direct' TF
relation from a cluster subject to observational selection effects by the
application of straightforward iterative procedure -- thus solving what has
been termed in the literature as the `calibration problem' (c.f. Willick 1994;
Hendry et al, 1995). It is important to recognise, however, that the
corresponding `raw' log distance estimator will {\em still\/} be biased,
in the sense of equation (4), at all true distances if applied to a galaxy
survey subject to magnitude selection effects. This is because the joint
distribution of absolute magnitude and log rotation velocity for
observable galaxies will {\em not\/} be equal to the intrinsic joint
distribution.
\\
Why has the use of the `direct' TF relation in recent literature continued
to be widespread? To understand the reason for this we must first note that
most recent analyses of galaxy distances and peculiar velocities have been
carried out within a Bayesian framework, thus involving the application
of Malmquist corrections to the `raw' log distance estimator. The
motivation for adopting a Bayesian approach (even if the Bayesian nature
of the problem has not always been explicitly acknowledged by authors!)
comes about from the way in which galaxy distance estimates and redshifts
have been combined in the majority of analyses. In both the early `toy'
parametric velocity field models of e.g. Lynden-Bell et al (1988),
Dressler \& Faber (1990), and the more sophisticated
reconstruction methods such as POTENT (c.f. Dekel et al 1990), essentially
galaxies are binned and grouped together and assigned radial peculiar
velocity estimates on the basis of their {\em estimated\/} distance.
The galaxy's actual distance could be radically different, and will depend on
the true spatial distribution of galaxies and the exact nature of the
survey selection function. Clearly galaxies which have small estimated
distance are more likely to have been scattered down from larger true
distances, since a volume element of fixed solid angle increases in size
with true distance; close to the limit of the survey volume, however, this
might no longer be the case, as galaxies scattered from larger true distances
might be too faint to be included in the redshift survey. By requiring that on
average the actual radial coordinate of the galaxy be equal to its estimated
distance, one would also ensure that on average the correct
peculiar velocity would be ascribed to that galaxy's apparent position.
The estimator which satisfies this condition can be defined following the
Bayesian approach outlined for the simple illustrative example of section 2,
and it is straightforward to show that such a `Malmquist corrected'
distance estimator, ${\bf{\hat{r}}}_{\rm{bayes}}$, satisfies the equation
\begin{equation}
{\bf{\hat{r}}}_{\rm{bayes}} = C \, \int \! {\rm{dexp}}
({\bf{\hat{w}}}) p({\bf{\hat{w}}}|{\bf{w}}) \, p({\bf{w}}) \, d{\bf{w}}
\end{equation}
where $C$ is a normalisation constant.
\\
The key point about equation (23) is that -- as before -- the
Bayesian distance estimator
depends upon the prior distribution of true log distance, $p({\bf{w}})$.
There has been no consensus in the literature on which prior one should
adopt. As we mentioned in section 2, in Lynden-Bell et al (1988) the prior
is assumed to correspond to a homogeneous distribution of galaxies --
thus defining homogeneous Malmquist corrections which are a function only
of distance. In Landy \& Szalay (1992), on the other hand, a more general
correction is derived by first estimating $p({\bf{w}})$ from
a spline fit to the histogram of log distance
{\em estimates\/} for the galaxies in the survey, thus in principle taking
into account inhomogeneities in the galaxy distribution. Due to the
sparseness of surveys, however, it is usually necessary to
average the distribution of galaxies over large solid angles, if not all, of
the sky. Therefore, the effects of clustering may still go largely
unaccounted for in the Landy \& Szalay prescription (c.f. Newsam et al, 1994).
In other recent analyses (c.f. Hudson 1994; Hudson et al 1995; Dekel 1994;
Willick 1994; Freudling et al 1995) a different method is proposed for
obtaining the prior distribution -- by
reconstructing the density field of optical or IRAS-selected galaxies
based on redshifts alone, assuming linear or mildly non-linear theory to
adequately describe the gravitational collapse of structure -- smoothed on a
scale of the order of 10 Mpc.
\\
In {\em all\/} of the above analyses the Malmquist corrections are derived
assuming that the conditional distribution of the `raw' log distance
estimator, $p({\bf{\hat{w}}}|{\bf{w}})$, is normally distributed at all
true log distances. As shown in Hendry \& Simmons (1994), this assumption
is invalid when the `raw' estimator is derived from the `Direct' TF
relation. Thus, the formula of Landy \& Szalay will result in an
{\em incorrect\/} Malmquist correction due to the bias of the `Direct'
TF log distance estimator. In general, if the prior distribution of true
log distance is inferred from the
observed distribution of log distance {\em estimates\/}, then one must
apply the formula of Landy \& Szalay using the `Inverse' TF estimator --
which we have seen {\em is\/} normally distributed and unbiased at all
true log distances, subject to the conditions specified above and in
Hendry \& Simmons (1994). A similar conclusion is reached in
Teerikorpi (1993), Feast (1994) and Freudling et al (1995).
\\
It is further shown in Hendry \& Simmons (1994) that the use of the `Direct'
TF relation as the raw log distance estimator in defining general
Malmquist corrections can only be justified if the prior distribution in
equation (23) corresponds to the {\em intrinsic\/} distribution of true log
distance. As a special case of this result, note that the homogeneous
Malmquist correction of Lynden-Bell et al (1988) applied to the
`Direct' TF estimator will therefore be valid provided that the intrinsic
distribution of galaxies is homogeneous. In a similar way, the
inhomogeneous corrections derived in Hudson et al (1995), Dekel (1994) and
Freudling et al (1995), will be valid provided the density field
reconstructed from optical or IRAS-selected surveys corresponds to the
intrinsic distribution of true log distance for the TF galaxies -- in other
words that the selection function of the redshift survey has been adequately
corrected for, and the redshift survey faithfully traces the {\em same\/}
underlying population as the galaxies to which the TF relation is being
applied.
\\
\section{\normalsize{CONCLUSIONS}}
In this paper we have set out to describe -- within the framework of
mathematical statistics -- some of the properties of `optimal' galaxy distance
estimators, including unbiasedness, sufficiency and efficiency. We have
shown that the intrinsic scatter of indicators
such as the Tully-Fisher and $D_{\rm{n}}-\sigma$ relations is sufficiently
large that the question of which statistical philosophy one should adopt
in the analysis of redshift surveys is far from trivial.
In particular we have seen that one may formulate the problem of galaxy
distance estimation as a Bayesian inference problem -- essentially the
approach which has been adopted implicitly in the literature in
defining Malmquist-corrected distance estimators -- but that there is no
general agreement over the issue of how one should then
best make use of prior information on the distribution of true galaxy
distances. In particular, a failure to adequately understand the
properties of the `raw' galaxy distance estimator used can lead to the
definition of invalid Malmquist corrections, as was the case in e.g.
Landy \& Szalay (1992). In Newsam, Simmons \& Hendry
(1995) we show that the use of such invalid corrections can frequently be
worse than applying no corrections at all. A similar conclusion was
reported in Freudling et al (1995), where it was shown that a number of
biases may have gone unresolved in earlier attempts to incorporate prior
information in the definition of distance estimators.
\\
In reality, the issue of defining an `optimal' galaxy distance estimator
is only the first part of the story. In applying the POTENT procedure, for
example, whether or not a distance estimator is biased is not the
crucial question; what is important is to construct an unbiased smoothed
peculiar velocity field. Although there appears some justification as to
why this procedure requires the application of an essentially Bayesian
approach, the Malmquist corrections which this approach entails
are strictly only valid if galaxies are not too sparse, the
gradient of the velocity field is not too large, and the effective radius
of the window function used to smooth the data is not too wide. In
Newsam, Simmons \& Hendry (1995) a Monte-Carlo procedure, involving
the generation of large numbers of `mock' redshift surveys, is devised and
implemented with the purpose of eliminating {\em all\/} biases from the
POTENT-recovered velocity and density fields -- not only those associated
with the scatter of the distance indicators. A similar algorithm may be
adopted for other reconstruction methods, and has the distinct advantage
of being easily adapted to more general (and more realistic!) selection
functions and distance indicators -- involving, e.g., correlations between
three or more observables where a wholly analytic treatment can often be
intractable (c.f. Hendry \& Simmons, 1994). A very similar Monte-Carlo
approach has been adopted in
Freudling et al (1995). These papers serve as an important reminder
that the question of galaxy distance estimation cannot be regarded in
isolation: ultimately the choice of which distance estimator is `optimal'
depends on the context in which the distance estimator is being used.
\\
It is perhaps worthwhile to end on a positive note. The use of redshift
independent galaxy distance indicators in conjunction with redshift surveys
has opened up an exciting -- and highly productive -- `industry' in
cosmology during the past decade or so. Although the statistical problems
arising from the large intrinsic scatter of these indicators are considerable,
the mathematical machinery briefly sketched in this paper equips us with the
necessary tools to address important issues such as their sensitivity to the
choice of prior information. Moreover, the significant recent advances made in
developing and applying more accurate distance indicators, such as surface
brightness fluctuations and the supernova light curve shape method, offer
some further cause for optimism: perhaps within the next decade we will be
able to map the large scale structure of the local universe with sufficient
accuracy that the question of whether one should adopt a Bayesian approach
to the analysis -- and how in detail it should be implemented -- will no
longer be important.
\\
Putting this another way, in the general statistics literature on
Bayesian inference, when one's results are sensitive to the choice of prior
information, one is usually advised to go out in search of better data.
Fortunately for those cosmologists measuring galaxy distances, it appears
that such data are indeed on their way!
\section*{\normalsize{REFERENCES}}
\footnotesize
\begin{itemize}
\item{Bertschinger E., Dekel A., 1989, Ap.J. (Lett), {\bf{336}}, L5}
\item{Chamcham K., Hendry M.A., 1995, MNRAS, submitted}
\item{Dekel A., 1994, Ann. Rev. Astr. Astrophys., 371 }
\item{Dekel A., Bertschinger E., Faber S.M., 1990, Ap.J.,
{\bf{364}}, 349}
\item{Dekel A., Bertschinger E., Yahil A., Strauss M.S.,
Davis M., Huchra J., 1993, Ap.J., {\bf{412}}, 1}
\item{Dressler A., 1994, in
`Cosmic velocity fields', eds. Lachieze-Rey M., Bouchet F., (Editions
Frontieres), 9}
\item{Dressler A., Faber S., 1990, Ap.J., {\bf{354}}, 13}
\item{Eastman R.G., Kirschner R.P, 1989, Ap.J., {\bf{347}}, 771}
\item{Eddington A., 1913, MNRAS, {\bf{73}}, 359}
\item{Faber S.M., Jackson R.E., 1976, Ap.J., {\bf{204}}, 668}
\item{Feast M., 1994, MNRAS, {\bf{266}}, 255}
\item{Federspiel M., Sandage A., Tammann G.A., 1994, Ap.J., {\bf{430}}, 29}
\item{Fiegelson E., Babul C.J., 1992, `Statistical Challenges in Modern
Astronomy', (Springer-Verlag)}
\item{Fisher K., Davis M., Strauss M., Yahil A., Huchra J.P.,
1994, MNRAS, {\bf{267}}, 927}
\item{Freedman W.L., et al, 1994, Nature, {\bf{371}}, 757}
\item{Freudling W., Da Costa L., Wegner G., Giovanelli R., Haynes M.,
Salzer J., 1995, A.J., in press}
\item{Giraud E., 1987, Astron. Astrophys., {\bf{174}}, 23}
\item{Gould A., 1995, Ap.J., {\bf{440}}, 510}
\item{Hamuy M., Phillips M.M., Maza J., Suntzeff N., Schommer R.,
Aviles R., 1995, A.J., {\bf{109}}, 1}
\item{Heavens, A.F., Taylor, A.N., 1995, MNRAS, in press}
\item{Hendry M.A., 1992a, PhD Thesis, Univ. of Glasgow}
\item{Hendry M.A., 1992b, Vistas in Astronomy, {\bf{35}}, 239}
\item{Hendry M.A., Rauzy S., Salucci P., Persic M., 1995,
Astrophys. Lett. \& Comm., in press}
\item{Hendry M.A., Simmons J.F.L., 1990, Astron. Astrophys.,
{\bf{237}}, 275}
\item{Hendry M.A., Simmons J.F.L., 1994, Ap.J., {\bf{435}}, 515}
\item{Hoel P., 1962, `An introduction to mathematical statistics',
(Wiley, NY)}
\item{Hogg R., Craig A., 1978, `An introduction to mathematical
statistics', (MacMillan)}
\item{Hudson M.J., 1994, MNRAS, {\bf{266}}, 468}
\item{Hudson M.J., Dekel A., Courteau S., Faber S.M., Willick J.A.,
1995, MNRAS, in press}
\item{Jacoby G.H., et al, 1992, PASP, {\bf{104}}, 599}
\item{Kanbur S.M., Hendry M.A., 1995, Astron. Astrophys., in press}
\item{Kendall M., Stuart A., 1963, `The advanced theory of statistics',
(Haffner, NY)}
\item{Kirschner R.P., Kwan J., 1974, Ap.J., {\bf{193}}, 27}
\item{Lahav O., Fisher K.B., Hoffman Y., Scharf C., Zaroubi S.,
1994, Ap.J. (Lett), {\bf{423}}, L93}
\item{Landy S., Szalay A., 1992, Ap.J., {\bf{391}}, 494}
\item{Lauer T.R., Postman M., 1994, Ap.J., {\bf{425}}, 418}
\item{Lynden-Bell D., Faber S.M., Burstein D., Davies R.D., Dressler A.,
Terlevich R.J., Wegner G., 1988, Ap.J., {\bf{326}}, 19}
\item{Malmquist K.G., 1920, `Medd. Lund. Astron. Obs.', 20}
\item{Malmquist K.G., 1922, `Medd. Lund.' Ser. I., 100}
\item{Mathewson D.S., Ford V.L., Buchhorn M., 1992, Ap.J. (Lett),
{\bf{389}}, L5}
\item{von Mises R., 1957, `Probability, statistics \& truth', (Allen \&
Unwin)}
\item{Mood A.M., Graybill A.F., 1974, `Introduction to the theory of
statistics', (McGraw-Hill, NY)}
\item{Mould J., Akeson R.L., Bothun G.D., Han M., Huchra J.P., Roth J.,
Schommer R.A., 1993, Ap.J., {\bf{409}}, 14}
\item{Newsam A.M., Simmons J.F.L., Hendry M.A., 1994, in
`Cosmic velocity fields', eds. Lachieze-Rey M., Bouchet F., (Editions
Frontieres), 49}
\item{Newsam A.M., Simmons J.F.L., Hendry M.A., 1995, Astron.
Astrophys., {\bf{294}}, 627}
\item{Nusser A., Davis M., 1994, Ap.J. Lett., {\bf{421}}, L1}
\item{Pierce M.J., Tully R.B., 1988, Ap.J., {\bf{330}}, 579}
\item{Pierce M.J., et al, 1994, Nature, {\bf{371}}, 385}
\item{Rauzy S., Lachieze-Rey M., Henriksen R.N., 1993,
Astron. Astrophys., {\bf{273}}, 357}
\item{Riess A.G., Press W.H., Kirschner R.P., 1995a, Ap.J. Lett.
{\bf{438}}, L17}
\item{Riess A.G., Press W.H., Kirschner R.P., 1995a, Ap.J. Lett.
{\bf{445}}, L91}
\item{Rubin V.C., Thonnard N., Ford W.K., Roberts M.S., 1976,
Astron. J., {\bf{81}}, 719}
\item{Saha A., Labhart L., Schwengeler H., Machetto F.D., Panagia N.,
Sandage A., Tammann G.A., 1994, Ap.J., {\bf{425}}, 14}
\item{Saha A., Sandage A., Labhart L.,Schwengeler H., Tammann G.A.,
Panagia N., Machetto F.D., 1995, Ap.J., {\bf{438}}, 8}
\item{Salucci P., Frenk C.S., Persic M., 1993, MNRAS, {\bf{262}}, 392}
\item{Sandage A., 1994, Ap.J., {\bf{430}}, 13}
\item{Sandage A., Tammann G., 1975a, Ap.J., {\bf{196}}, 313}
\item{Sandage A., Tammann G., 1975b, Ap.J., {\bf{197}}, 265}
\item{Sandage A., Tammann G., 1993, Ap.J., {\bf{415}}, 1}
\item{Schechter P.L., 1980, Astron.J., {\bf{85}}, 801}
\item{Schmidt B.P., Kirschner R.P. Eastman R.G., 1992, Ap.J., {\bf{395}},
366}
\item{Schmidt B.P., et al, 1994, Ap.J., {\bf{432}}, 42}
\item{Simmons J.F.L., Newsam A.M., Hendry M.A., 1995, Astron. Astrophys.,
{\bf{293}}, 13}
\item{Shaya E.J., Tully R.B., Pierce M.J., 1992, Ap.J., {\bf{391}}, 16}
\item{Shaya E.J., Tully R.B., Peebles P.J.E., 1995, Ap.J., in press}
\item{Sodre Jr.L., Lahav O., 1993, MNRAS, {\bf{260}}, 285}
\item{Teerikorpi P., 1993, Astron. Astrophys., {\bf{280}}, 443}
\item{Tonry J.L., Schneider D.P., 1988, A.J., {\bf{96}},807}
\item{Triay R., Lachieze-Rey M., Rauzy S., 1994, Astron. Astrophys.,
{\bf{289}}, 19}
\item{Tully R.B., Fisher J.R., 1977, Astron. Astrophys.,
{\bf{54}}, 661}
\item{van den Bergh, S., 1995, JRASC, {\bf{89}}, 6}
\item{Vaughan T.E., Branch D., Miller D.L., Perlmutter S.,
1995, Ap.J., {\bf{439}}, 558}
\item{Visvanathan N., Sandage A., 1977, Ap.J., {\bf{216}}, 214}
\item{Willick J., 1994, Ap.J. Supp., {\bf{92}}, 1}
\item{Zaccheo T.S., Gonsalves R.A., Ebstein S.M., Nisenson P.,
1995, Ap.J. Lett., {\bf{439}}, L43}
\end{itemize}
\end{document}
|
\section{Introduction}
In a recent paper [1] we showed that the theory of the strong
perturbations in
quantum mechanics [2] gives, at the leading order, an adiabatic behavior
for
the quantum system which is applied to. Beside, we showed, through the
interaction picture, that the validity condition $<H_0> \ll <V>$, being
the average taken on each eigenstate of the perturbation,
should be verified at any time. Actually, this is a set of conditions that
could give rise to secularities as we will show, being a secularity
a polynomial contribution in time in the perturbation term. The
secularities
are normally attributed to phase-mixing in the solution of the
Schr\"{o}dinger equation. Actually, a resummation can be
easily accomplished by comparing the result of the leading order of the
strong perturbation theory with the same in the interaction picture. The
only
effect, however, can be a harmless shift in the energy-levels of the
perturbation.
We can prove all that by deriving the unitary evolution operator for the
leading order, that is, the adiabatic evolution operator for the given
equation. In this way, the equations of higher orders in ref.[2] can be
integrated and the above cause of secularities pointed out. This problem
is exemplified through a two-level model that, however, still shows a
secularity at second order. However, it should be stressed that the
model here considered suffers similar problems from the standard small
perturbation theory too.
The full theory is applied to a quantum version of a well-known classical
model [3], that is a particle under the effect of two waves having
different amplitudes, frequencies and wave-numbers. We assume the
perturbation acting from the far past, being the particle free in that
limit. The model appears treatable by our method as is the one in ref.[4].
The paper is so structured. In sec.2 we derive the unitary evolution
operator and the higher order equations are solved. A possible origin of
secularities is given. In sec.3 we apply
the theory to the case of a two-level system in a constant perturbation
showing how secularities can be partially eliminated,
being one of the causes the one pointed out in sec.2.
In sec.4 we consider the quantized version of the classical model of the
particle in interaction with two waves.
In sec.5 some conclusions are drawn and the problem of the limitations of
the
theory is considered.
\section{Higher Order Terms and Secularities}
The general theory of strong pertrubations in quantum mechanics as
developed in ref.[2] considers a unitary evolution operator $U(t,t_0)$
such that
\begin{equation}
VU(t,t_0) = i\hbar\frac{dU(t,t_0)}{dt} \label{eq:main}
\end{equation}
being $V$ the perturbation. The general hamiltonian of the system has
the form $H=H_0 + V$ with $H_0$ that may also depend on time,
so that a perturbation series could be derived as
\begin{eqnarray}
|\psi(t)> &=& U|\psi(t_0)> \nonumber \\
&-& \frac{i}{\hbar}U\int_{t_0}^{t}dt'U^\dagger(t') H_0 U(t')
|\psi(t_0)>
\label{eq:series} \\
&+& \left(-\frac{i}{\hbar}\right)^2
U\int_{t_0}^{t}dt'U^\dagger(t') H_0 U(t')
\int_{t_0}^{t'}dt''U^\dagger(t'') H_0 U(t'')
|\psi(t_0)> \nonumber \\
&+& \cdots \nonumber
\end{eqnarray}
or, introducing the time ordering operator $T$, as
\begin{eqnarray}
|\psi(t)> &=& U|\psi(t_0)> \nonumber \\
&-& \frac{i}{\hbar}U\int_{t_0}^{t}dt'
U^\dagger(t') H_0 U(t')|\psi(t_0)>
\\ \label{eq:series_o}
&+& \frac{1}{2}
\left(-\frac{i}{\hbar}\right)^2U\int_{t_0}^{t}dt'
\int_{t_0}^{t}dt''
T(U^\dagger(t') H_0 U(t')
U^\dagger(t'') H_0 U(t''))|\psi(t_0)> \nonumber \\
&+& \cdots \nonumber
\end{eqnarray}
It is easy to see that the main problem is the determination of the
operator $U$ and then the computation of $U^\dagger H_0 U$ to go to
higher orders. This can be easily accomplished if we consider the main
results of ref.[1], that is, the leading order wave function,
$|\psi^{(0)}(t)> = U|\psi(t_0)>$, is just the following adiabatic one
\begin{equation}
|\psi(t)> \sim \sum_n c_n e^{i\gamma_n}
e^{-\frac{i}{\hbar}\int_{t_0}^t dt'v_n(t')}
|n;t>
\end{equation}
being $c_n = <n;t_0|\psi(t_0)>$, $V|n; t> = v_n(t)|n; t>$ and
\begin{equation}
\gamma_n(t) = \int_{t_0}^t dt' <n; t'|i\frac{d}{dt'}|n; t'>.
\end{equation}
This should
be compared with the one obtained in the interaction picture that gives
\begin{equation}
|\psi(t)> \sim \sum_n c_n e^{i\gamma_n}
e^{-\frac{i}{\hbar}\int_{t_0}^t dt'<n; t'|H_0|n,
t'>}
e^{-\frac{i}{\hbar}\int_{t_0}^t dt'v_n(t')} |n;t>
\label{eq:inter}
\end{equation}
where level shifts appear due to the unperturbed part of the hamiltonian.
We now show that these shifts could give rise to secularities for the
condition
$<n;t|H_0|n;t> \ll v_n(t)$.
Let us consider the evolution operator $U$ for an adiabatic evolution. It
is
a simple matter to see that we can write for eq.(\ref{eq:main})
\begin{equation}
U(t, t_0) = \sum_n e^{i\gamma_n(t)}
e^{-\frac{i}{\hbar}\int_{t_0}^t dt'v_n(t')}
|n;t><n;t_0|
\end{equation}
then
\begin{eqnarray}
U^\dagger H_0 U &=& \sum_n <n;t|H_0|n;t>|n;t_0><n;t_0| +
\sum_{m,n,m\neq n} e^{i[\gamma_n(t)-\gamma_m(t)]}
\\
& & e^{-\frac{i}{\hbar}\int_{t_0}^t
dt'[v_n(t')-v_m(t')]}
<m;t|H_0|n;t>|m;t_0><n;t_0|. \nonumber
\end{eqnarray}
We fix our attention on the first term on the rhs of the above equation.
That
term, when substitued in eq.(\ref{eq:series}), at the first order gives
\begin{eqnarray}
-\frac{i}{\hbar}U\int_{t_0}^t dt'
\sum_n <n;t'|H_0|n;t'>|n;t_0><n;t_0|\psi(t_0)> &=& \\
\sum_n \left(-\frac{i}{\hbar}\int_{t_0}^t dt'<n;t'|H_0|n;t'>\right)
c_n e^{i\gamma_n}
e^{-\frac{i}{\hbar}\int_{t_0}^t dt'v_n(t')}|n; t> & &
\nonumber
\end{eqnarray}
from which we recognize the second term of the series development of the
exponential of the level shifts in eq.(\ref{eq:inter}). This term could
give
rise to secularities in the perturbation series
if the shifts $<n; t|H_0|n; t>$ are time-independent as we
are going to show in the next section. It must be noticed that the above
term
is a direct application of the condition $<n;t|H_0|n;t> \ll v_n(t)$
and comes directly from the theory of strong perturbations.
So, as a rule, such terms should be simply resummed away.
This is accomplished without difficulty by computing the
level shifts and using eq.(\ref{eq:inter}) as leading order.
By comparing the level shifts with the energy levels of the
perturbation, or if the shifts are simply harmless,
we are able to realize if we can neglect such shifts. Otherwise,
we retain them and rewrite the evolution operator as
\begin{equation}
U(t, t_0) = \sum_n e^{i\gamma_n(t)}
e^{-\frac{i}{\hbar}\int_{t_0}^t dt'<n; t'|H_0|n;
t'>}
e^{-\frac{i}{\hbar}\int_{t_0}^t dt'v_n(t')}
|n;t><n;t_0|
\end{equation}
redefining the full perturbation series. We will clarify the above
arguments
with the following example.
\section{Two-Level System with a Constant Perturbation}
We consider the hamiltonian
\begin{equation}
H = H_0 + V = \left( \begin{array}{cc}
E_1 & 0 \\
0 & E_2
\end{array} \right) +
\left( \begin{array}{cc}
0 & V_{12} \\
V_{21} & 0
\end{array} \right)
\end{equation}
whose exact solution is well-known [5]. We apply to it the above results
to
exemplify the method.
The eigenstates of the perturbations are
\begin{equation}
|v_1> = \frac{1}{\sqrt{2}}\left( \begin{array}{c}
1 \\
-\frac{V_{21}}{|V_{12}|}
\end{array} \right),
|v_2> = \frac{1}{\sqrt{2}}\left( \begin{array}{c}
\frac{V_{12}}{|V_{12}|} \\
1
\end{array} \right)
\end{equation}
corresponding to the eigenvalue $-|V_{12}|$ and $|V_{12}|$ respectively.
Then, we have
\begin{equation}
U(t) = e^{\frac{i}{\hbar}|V_{12}|t}|v_1><v_1| +
e^{-\frac{i}{\hbar}|V_{12}|t}|v_2><v_2|
\end{equation}
and
\begin{eqnarray}
U^\dagger(t)H_0U(t) &=& \frac{E_1 + E_2}{2}I + \label{eq:uh0u} \\
& & \frac{E_1-E_2}{2|V_{12}|}
\left(V_{12}e^{-2\frac{i}{\hbar}|V_{12}|t}|v_1><v_2|+
V_{21}e^{2\frac{i}{\hbar}|V_{12}|t}|v_2><v_1|\right) \nonumber
\end{eqnarray}
being
\begin{equation}
<v_1|H_0|v_1> = <v_2|H_0|v_2> = \frac{E_1 + E_2}{2}.
\end{equation}
So, the first term in the rhs of eq.(\ref{eq:uh0u}) is just the
contribution
from the level shifts that, by comparing with eq.(\ref{eq:inter}),
reduces simply to a harmless phase factor and can be systematically
neglected, the condition $|V_{12}| \gg \frac{E_1 + E_2}{2}$
to be compared with the exact solution of this problem.
Our method works till second order as does the standard small perturbation
theory as, at that order, a term increasing with time appears. So, we have
found a possible cause of secularities but the problem is still open.
However,
it should be stressed that secularities are a general problem also for the
standard small perturbation theory [6], the question is to understand
the origin of them. This does not mean at all that the method is unuseful
as
we are going to show in the next section.
\section{The Two-Wave Model}
We consider the hamiltonian
\begin{equation}
H = \frac{p^2}{2m} + V_1 cos(k_1 x - \omega_1 t) +
V_2 cos(k_2 x - \omega_2 t)
\end{equation}
that can be easily quantized with the substitution
$p\rightarrow {\displaystyle -i\hbar\frac{\partial}{\partial x}}$.
This problem is analog
to the classical one in ref.[3] of a pendulum under the effect of an
oscillatory perturbation. We assume an adiabatic switching of the
perturbation from $t = -\infty$, being the particle free in
that limit, that is $\psi(x, -\infty) = \frac{1}{\sqrt{2\pi\hbar}}
e^{i\frac{p}{\hbar}x}$. This class of problems is tractable from the
point of view of secularities as already shown in ref.[4] and as we
are going to see in this case.
The leading order solution is then
\begin{eqnarray}
\psi^{(0)}(x,t) &=&\sum_{m=-\infty}^{+\infty}
\sum_{n=-\infty}^{+\infty}
J_m\left(\frac{V_1}{\hbar\omega_1}\right)
J_n\left(\frac{V_2}{\hbar\omega_2}\right) \times
\nonumber \\
& &e^{-i(mk_1+nk_2)x}e^{i(m\omega_1+n\omega_2)t}
\times \\
& &\frac{1}{\sqrt{2\pi\hbar}}e^{i\frac{p}{\hbar}x}
\nonumber
\end{eqnarray}
being $J_m$ and $J_n$ the Bessel functions of order $m$ and $n$
respectively.
It is quite simple to derive
\begin{eqnarray}
U^\dagger H_0 U \psi(x, -\infty) &=& \left(
\frac{p^2}{2m} + \frac{k_1^2 V_1^2}{4m\hbar^2\omega_1^2} +
\frac{k_2^2 V_2^2}{4m\hbar^2\omega_2^2}\right) \psi(x,-\infty) +
\nonumber \\
& &\sum_{\stackrel{m,r}{m\neq r}}\sum_{\stackrel{n,s}{n\neq s}}
J_m\left(\frac{V_1}{\hbar\omega_1}\right)
J_r\left(\frac{V_1}{\hbar\omega_1}\right)
J_n\left(\frac{V_2}{\hbar\omega_2}\right)
J_s\left(\frac{V_2}{\hbar\omega_2}\right) \times \label{eq:J}\\
& &\frac{(p-m\hbar k_1 -n\hbar k_2)^2}{2m} \times \nonumber \\
& &e^{-i[(m-r)k_1+(n-s)k_2)]x}e^{i[(m-r)\omega_1+(n-s)\omega_2)]t}
\psi(x,-\infty) \nonumber
\end{eqnarray}
having put
$H_0 = {\displaystyle -\frac{\hbar^2}{2m}\frac{\partial^2}{\partial
x^2}}$.
It easy to
see the part originating the secularities. This is the first term on the
rhs
that gives rise to a phase-factor due to the kinetic energy of the
particle
and the ponderomotive forces of the two waves. This term can be summed
away.
The sums in eq.(\ref{eq:J}) can be evaluated and this yields the
development
parameter of the perturbation series. We avoid this calculation here being
not the main point of the paper, we just point out that no secularities
appear at this order.
\section{Conclusions}
With the fundamental result of ref.[1], the theory of strong perturbations
in quantum mechanics is strictly linked with the adiabatic theorem. This
means that all the limitations of the adiabatic theorem should be ported
to
this theory. One problem may be due to a continuos spectrum of eigenvalues
for the perturbation. We have avoided to face this problem in sec.4,
although meaningful results was obtained.
However, the main question remains the secularities arising from
the perturbative solution of the Schr\"{o}dinger equation. We are not
assured that going to higher orders, secularities will not appear. We
would
like to stress again that this kind of problems arise normally in the
standard theory of small perturbations as could be seen in ref.[6] where
some methods are given to face the question.
Anyway, we showed with a lot of examples that the theory is indeed
useful and a wide possibility to explore new solutions of the
Schr\"{o}dinger equation is surely open. This in turn means that new
quantum behaviors should be considered, that is, the class of quantum
systems
that we define strongly perturbed.
\newpage
[1] M.Frasca, ``{\sl The Leading Order of the Theory of Strong
Perturbations in Quantum Mechanics}'', quant-ph/9507007 (1995);
M.Frasca, submitted to Nuovo Cimento (1995);
see also A.Joye, Ann. Inst. Henri Poincare', {\bf 63}, 231 (1995)
[2] M.Frasca, Phys. Rev. A {\bf 45}, 43 (1992);
Phys. Rev. A {\bf 47}, 2364 (1993)
[3] R.Z.Sagdeev, D.A.Usikov, G.M.Zaslavsky,
{\sl Nonlinear Physics (From the Pendulum to Turbulence and Chaos)},
Harwood Academic Publishers, Philadelphia, Penn., 1992
[4] M.Frasca, Nuovo Cimento B, {\bf 107}, 845 (1992);
Nuovo Cimento B, {\bf 109}, 1227 (1994); Phys. Rev. E, {\bf 53}, 1236
(1996)
[5] J.J.Sakurai, {\sl Modern Quantum Mechanics}, Benjamin/Cummings
Publishing Company, Reading, Mass., 1985
[6] J.Kevorkian, J.D.Cole, {\sl Perturbation Methods in Applied
Mathematics}, Springer-Verlag, New York, N.Y., 1985
\end{document}
|
\section{Introduction}{\label{S.1}}
Because of their conceptual simplicity, path-integral methods \ct{ref1,pol}
often provide convenient procedures to obtain insight in field theoretical
problems. In recent work by Strassler, McKeon, Schmidt, Schubert and others
\ct{strass}-\ct{vn1} world-line path integrals have been applied to a
reformulation of standard Feynman perturbation theory from which useful
information on the structure of perturbative Green's functions is extracted.
Some of these results were actually first derived in the particle-limit of
string theory \ct{bern}.
A basic question in this context is the representation of propagators in
quantum field theory by path integrals for relativistic particles of various
kind. In particular one would like to know the classical actions to be used
in these path-integrals, as well as the precise meaning of the functional
measure in some regularized form, e.g.\ by discretization. Answers to these
questions establish firm connections between the so-called first- and second
quantized formulations of relativistic quantum theory.
This paper addresses both of these problems. It pursues them from two
complementary points of view. First of all, there is the pragmatic question
of how to convert a given field-theory propagator to a path integral
expression. Such a procedure is discussed in this paper for both spin-0 and
spin-1/2 particles. Starting at the other end, one can also ask what kind of
quantum field theory is associated with a given classical action. The kind
of actions considered in this context are usually taken to possess some
desirable properties like reparametrization invariance and supersymmetry,
which pose however additional difficulties to quantization procedures since
gauge fixing then becomes necessary. Using previously established
BRST-procedures \ct{jw1,jw3} this is done in an extended state space with a
number of ghost- and auxiliary degrees of freedom. The propagators in these
extended state spaces are obtained, and it is then shown how to reduce them to
standard expressions in terms of physical variables only.
One result which deserves to be emphasized is that the irreducible
path-integral expression for the propagator of a massive Dirac fermion
is not based on a manifestly supersymmetric action, although
world-line supersymmetry is realized algebraically in the quantum theory
by the Dirac operator. Indeed, it is shown that the manifestly supersymmetric
version of the theory contains a `hidden' topological fermionic degree
of freedom which doubles the number of components of the physical states.
This theory therefore describes a degenerate doublet of fermions, rather
than a single massive Dirac particle. A path-integral expression for a
single massive Dirac fermion is also obtained (sect.\ref{S.7}). As in other
cases, the difference between the two models can be traced to the
representation of $\gam_5$.
This paper is organized as follows. In sects.(\ref{S.2})-(\ref{S.6}) the
free field theory of a scalar spin-0 particle is considered. A simple and
well-known path-integral expression is obtained, which is subsequently
rederived from the reparametrization-invariant classical model. These
calculations also help to explain the general procedures used.
In sect.(\ref{S.7}) a path-integral for a spin-1/2 particle is derived.
For massless particles it agrees with results in the literature \ct{cohen,hht}.
For massive fermions a new term is present in the action which has not
been considered before, and which is based on a bosonic representation of
$\gam_5$. The manifestly supersymmetric theory, which uses a fermionic
representation for $\gam_5$, is analyzed in sects.(\ref{S.8})-(\ref{S.10}),
and is shown to describe a doublet of spinors if the mass is non-zero.
Our conclusions are presented in sect.(\ref{S.11}).
\section{Free particle propagators}{\label{S.2}}
The Feynman propagator for a free scalar particle of mass $m$ in
$D$-dimensional space-time is a specific solution of the inhomogeneous
Klein-Gordon equation
\be
\lh - \Box_x + m^2 \rh\, \Del_F (x - y)\, =\, \del^D(x - y),
\label{2.0}
\ee
\nit
such that positive frequencies propagate forward in time, and negative
frequencies backward. An explicit expression in terms of a Fourier
integral is
\be
\Del_{F}(x - y)\, =\, \int \frac{d^D p}{(2\pi)^D}\,
\frac{e^{i p \cdot (x-y)}}{p^2 + m^2 - i \ve} .
\label{2.1}
\ee
\nit
In the limit $\ve \rightarrow 0^+$ the simple real pole at positive $p^0 =
E(\vec{p}) = \sqrt{\vec{p}^{\,2} + m^2}$ gives the mass of the particle as
$m = E(0)$. The arbitrarily small imaginary part $i\ve$
implements the causality condition.
There is a straightforward connection between this propagator and the
classical mechanics of a relativistic point-particle through the
path-integral formalism \ct{ref1}. To establish this, let us first consider
the very general problem of finding the inverse of a non-singular hermitean
operator $\hH$. Following Schwinger \ct{ref2}, we construct the formal
solution
\be
\hH^{-1}\, =\, \lim_{\ve \rightarrow 0^+}\,
i \int_0^{\infty} dT\, e^{-iT \lh \hH - i\ve \rh}.
\label{2.2}
\ee
\nit
Here the exponential operator
\be
\hK_{\ve}(T)\, =\, e^{-iT \lh \hH - i\ve \rh}
\label{2.3}
\ee
\nit
is the solution of the Schr\"{o}dinger equation
\be
i \dd{\hK_{\ve}(T)}{T}\, =\, \lh \hH - i\ve \rh\, \hK_{\ve}(T),
\label{2.4}
\ee
\nit
with the special properties\footnote{$\hI$ denotes the unit operator.}
\be
\hK_{\ve}(0) = \hI, \hspace{3em}
\lim_{T \rightarrow \infty} \hK_{\ve}(T) = 0.
\label{2.5}
\ee
\nit
If the operator $\hH$ acts on the single-particle state-space with a complete
co-ordinate basis $\left\{ |x\rangle \right\}$, the matrix elements of the
operator in the co-ordinate basis are
\be
K_{\ve}(x - y|T)\, =\, \langle x| \hK_{\ve}(T) | y \rangle .
\label{2.6}
\ee
\nit
Completeness of the basis then implies Huygens' composition principle
\be
\int d^D x^{\prime}\, K_{\ve^{\prime}}(x - x^{\prime}|T^{\prime})\,
K_{\ve^{\prime\prime}}(x^{\prime} - x^{\prime\prime}|T^{\prime\prime})\,
=\, K_{\ve}(x - x^{\prime\prime}|T^{\prime} + T^{\prime\prime}).
\label{2.10}
\ee
\nit
Note that $\ve = (\ve^{\prime} T^{\prime} + \ve^{\prime\prime}
T^{\prime\prime})/(T^{\prime} + T^{\prime\prime})$ stays arbitrarily small if
$\ve^{\prime}$ and $\ve^{\prime\prime}$ are small enough.
Repeated use of eq.(\ref{2.10}) now allows one to write
\be
K_{\ve}(x-y|T)\, =\, \int \prod_{n=1}^N d^Dx_n \prod_{m=0}^N
K_{\ve}\lh x_{m+1} - x_m | \Del T \rh,
\label{2.6.1}
\ee
\nit
with $\Del T = T/(N+1)$, and $x_0 = y$, $x_{N+1} = x$. Keeping $T$ fixed,
the limit $N \rightarrow \infty$ becomes an integral over continuous (but
generally non-differentiable) paths in co-ordinate space-time between points
$y$ and $x$. (Observe that $K_{\ve}(x-y|\Del T)$ depends only on the difference
$(x-y)$ and converges to $\del^D(x-y)$ for $\Del T \rightarrow 0$.)
If the operator $\hH$ is an ordered expression in terms of a canonical set of
operators $(\hx^{\mu},\hp_{\mu})$:
\be
\hH\, =\, \sum_{k,l}\, \hp_{\mu_1} ... \hp_{\mu_k} H^{\mu_1 ... \mu_k}_{\nu_1
... \nu_l} \hx^{\nu_1} ... \hx^{\nu_l},
\label{2.7}
\ee
\nit
then we can expand the co-ordinate path-integral expression (\ref{2.6.1})
further to a phase-space path-integral
\be
\ba{lll}
K_{\ve}(x - y|T) & = & \dsp{ \frac{1}{(2\pi)^D}\, \int d^D p_0
\int \prod_{n=1}^{N} \frac{d^D x_n d^D p_n}{(2\pi)^D}\,
e^{ i \sum_{k=0}^{N} \lh p_k \cdot (x_{k+1} - x_k)
- \Del T H(p_k, x_k) \rh} } \\
& & \\
& \rightarrow & \dsp{ \int_y^x \cD p(\tau) \cD x(\tau)\,
e^{ i \int_0^T d\tau \lh p \cdot \dot{x} - H(p,x) \rh }. }
\ea
\label{2.8}
\ee
\nit
Here $H(p,x)$ is the c-number symbol of the ordered operator $\hH$, and
we have tacitly assumed that the ordered symbol of the exponential can be
replaced by the exponential of the ordered symbol. This is certainly correct
for the main applications we consider in this paper, as may be checked by
explicit calculations.
It is now clear, that one may interpret the symbol $H(p,x)$ as the hamiltonian
of some classical system, and the argument of the exponential as the
classical action. Integration over the momentum variables $p(\tau)$ then
in general leads to the lagrangian form of this action
\be
K_{\ve}(x - y|T)\, =\, \int_y^x \cD x(\tau)\,
e^{i \int_0^T d\tau L(\dot{x}, x) },
\label{2.9}
\ee
\nit
where the precise meaning of the integration measure can be recovered
either from the phase-space expression (\ref{2.8}), or from requiring the
path-integral to satisfy Huygens' composition principle (\ref{2.10}).
Returning to eq.(\ref{2.0}), it states that $\Del_F(x - y)$ is the inverse
of the Klein-Gordon operator (in the space of square-integrable functions).
Rescaling it for later convenience by a factor $1/2m$, we consider the
evolution operator
\be
\hK_{\ve}(T)\, =\, \exp \left\{-\frac{iT}{2m} \lh -\Box + m^2 - i\ve \rh
\right\}.
\label{2.10.1}
\ee
\nit
In the co-ordinate representation the explicit expression for the
matrix element of this operator is
\be
K_{\ve}(x-y|T)\, =\, -i \lh \frac{m}{2 \pi T}\rh^{D/2}\,
e^{i \frac{m}{2T} (x-y)^2 - \frac{iT}{2} (m - i \ve)}.
\label{2.10.2}
\ee
\nit
The Feynman propagator can then be written as
\be
\Del_F(x-y)\, =\, \frac{i}{2m}\, \int_0^{\infty} dT K_{\ve}(x-y|T),
\label{2.10.3}
\ee
\nit
and using the previous results it can be cast in the form of a path integral
\ct{cas}
\be
\Del_F(x - y)\, =\, \frac{i}{2m}\, \int_0^{\infty} dT\, \int_y^x \cD
x(\tau)\, \exp \left\{\frac{im}{2}\,
\int_0^T d\tau \lh \dot{x}_{\mu}^2 - 1 \rh \right\}.
\label{2.11}
\ee
\nit
The same expression is also obtained directly by iteration of (\ref{2.10.2})
using the product formula (\ref{2.6.1}). For massless particles yet another
derivation, based on the concepts of moduli space, has been discussed in
\ct{cohen,hht}. The result shows, that the scalar propagator is connected
to a classical particle model with lagrangian
\be
L = \frac{m}{2}\, \lh \dot{x}_{\mu}^2 - 1 \rh ,
\label{2.11.1}
\ee
\nit
related by Legendre transform to the simple hamiltonian
\be
H(p,x)\, =\, \frac{p_{\mu}^2 + m^2}{2m}.
\label{2.12}
\ee
\nit
Because $H(p,x)$ is quadratic in $p_{\mu}$ and independend of $x^{\mu}$,
the integration measure in the path integral (\ref{2.11}) is just the free
gaussian measure of ref.\ct{ref1}.
The path-integral representation of the propagator establishes a simple
connection between scalar quantum field theory and the classical point
particle model (\ref{2.11.1}). As is well-known, this connection does not
only hold at the level of the action or hamiltonian, it also extends to the
dynamics, in the sense that paths in the neighborhood of the solutions of the
classical equations of motion give the dominant contribution to the path
integral, certainly for the simple model discussed where the path integral is
a pure Gaussian.
\section{Reparametrization invariance}{\label{S.3}}
The classical equations of motion which follow from the lagrangian
(\ref{2.11.1}) state that the momentum is constant along the
particle worldline:
\be
\dot{p}^{\mu} = m \ddot{x} = 0,
\label{3.1}
\ee
\nit
Therefore this classical theory seems to contain less information
than the quantum theory from which we started: we have to recover
the condition that the momentum should lie on the mass shell:
\be
p^2_{\mu} + m^2 = 0.
\label{3.2}
\ee
\nit
This condition is equivalent to the vanishing of the hamiltonian
(\ref{2.12}). Since the hamiltonian is the generator of translations
in the worldline parameter $\tau$, the mass-shell condition is
recovered by requiring the dynamics of the particle to be independent
of the world-line parametrization.
As is well-known \ct{brink1,brink2}, this can be achieved by introducing a
gauge variable $e(\tau)$ for re\-par\-am\-etri\-za\-tions of the world line
(the einbein). Under local reparametrizations $\tau \rightarrow \tau^{\prime}
= f(\tau)$, let the co-ordinates and einbein transform as
\be
x^{\mu}(\tau) \rightarrow x^{\prime \mu} (\tau^{\prime}) = x^{\mu}(\tau),
\hspace{3em}
e(\tau) \rightarrow e^{\prime}(\tau^{\prime})
= e(\tau) \frac{d\tau}{d\tau^{\prime}}.
\label{3.3}
\ee
\nit
Then a reparametrization-invariant action can be constructed of the form
\be
S_{cl}[x^{\mu}(\tau),e(\tau)]\, =\, \frac{m}{2}\, \int_0^T d\tau\,
\lh \frac{1}{e}\, \dot{x}_{\mu}^2 - e \rh.
\label{3.4}
\ee
\nit
The equations of motion one derives from this action are
\be
\frac{1}{e} \frac{d}{d\tau}\, \frac{1}{e} \frac{dx^{\mu}}{d\tau}\, =\, 0,
\hspace{3em} \lh \frac{1}{e} \frac{dx^{\mu}}{d\tau} \rh^2\, +\, 1\, =\, 0.
\label{3.5}
\ee
\nit
Identifying $d\tilde{\tau} = e d\tau$ with the proper time, and as a
consequence $p^{\mu} = m dx^{\mu}/d\tilde{\tau}$ with the proper momentum,
this reproduces both the world-line equation of motion and the mass-shell
condition.
For the new hamiltonian we can take
\be
H = \frac{e}{2m}\, \lh p_{\mu}^2 + m^2 \rh .
\label{3.6}
\ee
\nit
This generates proper-time translations on special phase-space functions
$F(x(\tau),p(\tau))$ by the Poisson-brackets:
\be
\frac{dF}{d\tilde{\tau}}\, =\, \frac{1}{e} \frac{dF}{d\tau}\,
=\, \frac{1}{e}\, \left\{ F, H \right\} .
\label{3.7}
\ee
\nit
Note, that one cannot impose the constraint that $H$ vanishes before
we compute the brackets.
The difficulty with this formulation is evidently the additional dynamical
variable $e(\tau)$, for which the evolution is not fixed by the Euler-Lagrange
equations derived from $S_{cl}[x^{\mu},e]$, and which has no conjugate
momentum. The origin of these difficulties is precisely the reparametrization
invariance (\ref{3.3}). As a result, the hamiltonian evolution equation
(\ref{3.7}) does not hold for arbitrary functions on the {\em complete} phase
space, $F(x^{\mu},p^{\mu},e)$. Clearly, to recover the results of
sect.\ref{S.2} it is necessary to fix $e(\tau)$ to a constant value and change
$\tau \rightarrow \tilde{\tau}$, which amounts to a rescaling of the unit of
time on the worldline such that the particle's internal clock and the
laboratory clock tick at the equal rates when the particle is at rest.
Therefore one chooses a gauge
\be
\dot{e} = 0,
\label{3.8}
\ee
\nit
implying that $e$ is constant, and adds this condition to the equations of
motion. But this can only be done after the variation of $e$ in the
action has produced the mass-shell constraint (\ref{3.2}). It is
therefore of interest to have a formalism in which one can impose the
gauge condition (\ref{3.8}) from the start, and still keep track of all the
constraints imposed by the reparametrization invariance (\ref{3.3}). An
appropriate formalism to solve this problem is the BRST procedure (for an
introduction, see for example refs.\ct{KO,henn,jw2}), which we describe here
for the case at hand \ct{jw1}.
\section{BRST formulation}{\label{S.4}}
We impose the gauge condition (\ref{3.8}) by adding it to the action with
a Lagrange multiplier $\lb$; at the same time we introduce a corresponding
Faddeev-Popov ghost action, using (real) anti-commuting ghost variables
$(b,c)$, in such a way that the complete action is invariant under the
Grassmann-odd, nilpotent BRST transformations
\be
\ba{llllll}
\dL x^{\mu} & = c \dot{x}^{\mu}, \hspace{2em} &
\dL e & = c \dot{e} + \dot{c} e, \hspace{2em} & \dL c & = c \dot{c}, \\
& & & & & \\
\dL b & = - i \lb, & \dL \lb & = 0.
\ea
\label{3.9}
\ee
\nit
After a convenient rescaling $ec \rightarrow c$, turning the ghost $c$ into a
world-line scalar density such that $\dL e = \dot{c}$ and $\dL c = 0$, the
BRST-invariant gauge fixed action reads
\be
S_{gf}\, =\, S_{cl}[x(\tau), e(\tau)]\, +\, \int_0^T d\tau\,
\lh \lb \dot{e} + i \dot{b} \dot{c} \rh.
\label{3.10}
\ee
\nit
Actually, because a partial integration has been performed in the ghost term,
the action is invariant only modulo a total time derivative. This is
sufficient. Note, that the Lagrange multiplier $\lb$ now plays the role of
momentum $p_e$ conjugate to the einbein $e$.
The action $S_{gf}$ has no local invariances left and can be treated by the
usual procedures of canonical hamiltonian analysis. The canonical momenta
are
\be
\ba{llll}
p^{\mu} & \dsp{ = \frac{m}{e}\, \dot{x}^{\mu}, } \hspace{2em}
& p_e & = \lb, \\
& & & \\
\pi_c & = - i \dot{b}, & \pi_b & = i \dot{c}.
\ea
\label{3.11}
\ee
\nit
Note that the ghost-momenta $(\pi_c, \pi_b)$ are imaginary.
The gauge-fixed hamiltonian is
\be
H_{gf}\, =\, \frac{e}{2m}\, \lh p_{\mu}^2 + m^2 \rh - i \pi_b \pi_c.
\label{3.12}
\ee
\nit
The equations of motion are given by the Poisson brackets
\be
\frac{dF}{d\tau}\, =\, \left\{F, H_{gf} \right\},
\label{3.13}
\ee
\nit
where the brackets are defined by
\be
\ba{lll}
\left\{F, G \right\} & = & \dsp{ \dd{F}{x^{\mu}} \dd{G}{p_{\mu}}\, -\,
\dd{F}{p_{\mu}} \dd{G}{x^{\mu}}\, +\, \dd{F}{e} \dd{G}{p_e}\, -\,
\dd{F}{p_e} \dd{G}{e} } \\
& & \\
& & \dsp{ +\, (-1)^{a_F} \lh \dd{F}{c} \dd{G}{\pi_c} + \dd{F}{\pi_c} \dd{G}{c}
\, +\, \dd{F}{b} \dd{G}{\pi_b} + \dd{F}{\pi_b} \dd{G}{b} \rh. }
\ea
\label{3.14}
\ee
\nit
Here all derivatives are taken from the left, and $a_F$ is the Grassmann parity
of $F$.
To recover the constraints imposed by local reparametrization invariance,
one constructs the (Grassmann-odd) conserved BRST charge
\be
\Og\, =\, \frac{c}{2m}\, \lh p_{\mu}^2 + m^2 \rh\, -\, i \pi_b p_e,
\hspace{3em} \left\{ \Og, H_{gf} \right\}\, =\, 0.
\label{3.15}
\ee
\nit
This BRST charge is nilpotent in the sense that
\be
\left\{ \Og, \Og \right\}\, =\, 0.
\label{3.15.1}
\ee
\nit
The BRST principle makes use of this property, and states that the physical
observables of the theory are the cohomology classes of the BRST operator
on the phase-space functions $F(x,p;e,p_e;c,\pi_c;b,\pi_b)$ with ghost
number\footnote{We assign ghost number +1 to $(c,\pi_b)$, and -1 to
$(b,\pi_c)$.} $N_{gh} = 0$ \ct{ff,henn2}. More precisely, physical
quantities are BRST invariant:
\be
\left\{ F, \Og \right\}\, =\, 0,
\label{3.16}
\ee
\nit
but this allows an ambiguity in $F$ as a result of (\ref{3.15.1});
this ambiguity is resolved by associating obervables $F$ with equivalence
classes of functions differing only by a BRST transformation:
\be
F^{\prime}\, \sim\, F \hspace{2em} \Leftrightarrow \hspace{2em}
F^{\prime}\, =\, F\, +\, \left\{ \Lb, \Og \right\},
\label{3.17}
\ee
\nit
with $\Lb$ a phase-space function of Grassmann parity opposite to $F$,
and one unit in ghost-number lower. In particular, any function which can be
written as a pure BRST transform
\be
G\, =\, \left\{ \Lb, \Og \right\},
\label{3.18}
\ee
\nit
is in the equivalence class of zero and may be taken to vanish. In the case
of the relativistic particle, this happens to be true for the hamiltonian
itself:
\be
H_{gf}\, =\, - \left\{ e\pi_c, \Og \right\},
\label{3.19}
\ee
\nit
and for the classical hamiltonian as well:
\be
p_{\mu}^2\, +\, m^2\, =\, -2m \left\{ \pi_c, \Og \right\}.
\label{3.20}
\ee
\nit
Hence the classical and ghost terms in the hamiltonian $H_{gf}$ vanish
separately and the mass-shell condition follows. With
\be
p_e\, =\, -i\, \left\{ b, \Og \right\},
\label{3.21}
\ee
\nit
we also recover the vanishing of the momentum conjugate to the einbein $e$.
Before turning to the quantum theory, we draw attention to a peculiarity of
the action (\ref{3.10}): it is invariant under SO(2) rotations in the ghost
variables $(b,c)$. Therefore this theory possesses a second BRST invariance
with conserved nilpotent charge
\be
\tilde{\Og}\, =\,\frac{b}{2m}\, \lh p_{\mu}^2 + m^2 \rh\, -\, i \pi_c p_e,
\hspace{3em} \left\{ \tilde{\Og}, H_{gf} \right\}\, =\, 0.
\label{3.22}
\ee
\nit
The algebra of these two BRST charges has the property that
\be
\left\{ \Og, \tilde{\Og} \right\}\, =\,
\frac{i}{m}\, \lh p_{\mu}^2 + m^2 \rh\, p_e.
\label{3.23}
\ee
\nit
Note that both factors on the right-hand side are separately conserved,
and vanish on the physical hyperplane in phase space.
\section{BRST quantization}{\label{S.5}}
Having formulated a complete, BRST-invariant (pseudo-)classical mechanics
for the relativistic point particle, we now return to the quantum theory and
study the relation between this model and quantum field theory. Various
aspects of this problem have been discussed in \ct{henn3,mon,jw1}. To obtain
additional insight, we construct a quantum theory corresponding to the
classical hamiltonian (\ref{3.12}) in an extended state-space, compute the
propagator and establish the relation with the usual Feynman propagator
(\ref{2.1}), thereby showing the physical equivalence of these different
formulations of scalar field theory.
The quantum theory of interest is obtained by replacing the
phase-space variables $(x^{\mu}, p_{\mu}; e, p_e; c, \pi_c; b, \pi_b)$ by
operators, and postulating (anti-)\-commutation relations between them in
direct correspondence with the Poisson-brackets (\ref{3.14}):
\be
\ba{llll}
\left[ \hx^{\mu}, \hp_{\nu} \right] & =\; i \del^{\mu}_{\nu}, &
\left[ \he, \hp_e \right] & =\; i, \\
& & & \\
\left\{ \hc, \hat{\pi}_c \right\} & =\; - i, &
\left\{ \hb, \hat{\pi}_b \right\} & =\; - i.
\ea
\label{5.1}
\ee
\nit
In the co-ordinate representation these algebraic relations hold on
making the identification
\be
\ba{llll}
\hp_{\mu} & \dsp{ =\, - i\dd{}{x^{\mu}},} & \hp_e & \dsp{ =\, - i\dd{}{e}, }\\
& & & \\
\hat{\pi}_c & \dsp{ =\, - i \dd{}{c},} & \hat{\pi}_b & \dsp{ =\, -i\dd{}{b}.}
\ea
\label{5.2}
\ee
\nit
The bosonic momenta are self adjoint, and the fermionic ones
anti self-adjoint with respect to the inner product
\be
\lh \Phi, \Psi \rh\, =\, i \int d^Dx^{\mu} de \int db dc\, \Phi^* \Psi,
\label{5.2.1}
\ee
\nit
where wave functions like $\Psi$ are polynomials in the ghost variables,
with co-efficients depending on the remaining co-ordinates $(x^{\mu},e)$:
\be
\Psi(b,c)\, =\, \psi - ib \psi_b + c \psi_c - icb \psi_{cb}.
\label{5.2.2}
\ee
\nit
The inner product (\ref{5.2.1}) then reads in components
\be
\lh \Phi, \Psi \rh\, =\, \phi^* \psi_{cb} + \phi^*_b \psi_c + \phi^*_c \psi_b
+ \phi^*_{cb} \psi.
\label{5.2.3}
\ee
\nit
Clearly this form is not positive definite. However, with respect to
this inner product the hamiltonian operator,
\be
\hH_{gf}\, =\, \frac{e}{2m}\, \lh - \Box + m^2 \rh\, +\,
i \frac{\pl^2}{\pl b\pl c},
\label{5.3}
\ee
\nit
and the nilpotent BRST operator,
\be
\hOg\, =\, \frac{c}{2m}\, \lh - \Box + m^2 \rh\, +\,
i \frac{\pl^2}{\pl e \pl b}, \hspace{3em}
\hOg^2\, =\, 0,
\label{5.4}
\ee
\nit
are self adjoint. The indefinite metric implicit in this inner product
is a general and necessary feature of a model with a nilpotent self-adjoint
operator like $\hOg$ \ct{jw3}.
It is also possible to define a positive-definite inner product on the
state space: introduce a duality operation \ct{jw6}
\be
\Psi\, \rightarrow\, \tilde{\Psi}\, =\, \psi_{cb} - ib \psi_c + c \psi_b
- icb \psi,
\label{5.5}
\ee
\nit
and define
\be
\ba{lll}
\langle \Phi, \Psi \rangle & = & \lh \Phi, \tilde{\Psi} \rh \\
& & \\
& = & \phi^* \psi + \phi^*_b \psi_b + \phi^*_c \psi_c + \phi^*_{cb} \psi_{cb}.
\ea
\label{5.6}
\ee
\nit
With respect to this positive definite inner product the fermionic momenta
and the BRST charge are no longer (anti-)self adjoint. In particular $\hc^*
= i \hat{\pi}_c$, $\hb^* = i \hat{\pi}_b$, and
\be
\langle \hOg \Phi, \Psi \rangle\, =\, \langle \Phi, \hOg^* \Psi \rangle,
\label{5.7}
\ee
\nit
where the adjoint BRST charge, also known as the co-BRST operator, is
\be
\hOg^*\, =\, \frac{1}{2m}\, \lh - \Box + m^2 \rh\, \dd{}{c}\,
+\, ib \dd{}{e}.
\label{5.8}
\ee
\nit
Like the BRST charge itself, $\hOg^*$ is nilpotent; however, it is not
conserved:
\be
\left[ \hOg^*, \hH_{gf} \right]\, =\, \tilde{\Og},
\label{5.9}
\ee
\nit
where $\tilde{\Og}$ is the operator corresponding to the dual BRST charge
we have encountered before in (\ref{3.22}):
\be
\tilde{\Og}\, =\,
\frac{ib}{2m} \lh -\Box + m^2 \rh\, +\, \frac{\pl^2}{\pl e \pl c}.
\label{5.10}
\ee
\nit
We also note in passing, that the BRST operator $\hOg$ itself is obtained
from the commutator of the hamiltonian with the adjoint of the
dual BRST charge:
\be
\left[ \tilde{\Og}^*, \hH_{gf} \right]\, =\, - \hOg,
\label{5.11}
\ee
\nit
with
\be
\tilde{\Og}^*\, =\, - \frac{i}{2m}\, \lh -\Box + m^2 \rh\, \dd{}{b}\,
-\, c \dd{}{e}.
\label{5.12}
\ee
\nit
We can now show how the physical states of the scalar particle are reobtained
in the BRST formalism through the cohomology of $\hOg$. One identifies the
physical states with equivalence classes of states which are BRST invariant
and differ only by a BRST-exact term:
\be
\hOg\, \Psi_{phys}\, =\, 0, \hspace{3em}
\Psi_{phys}\, \sim\, \Psi^{\prime}_{phys}\, =\,
\Psi_{phys}\, +\, \hOg\, \Lambda.
\label{5.13}
\ee
\nit
To obtain exactly one representative of each BRST invariance class,
consider the zero-modes of the BRST laplacian \ct{jw3,jw1}
\be
\ba{lll}
\Del_{BRST} & = & \hOg \hOg^*\, +\, \hOg^* \hOg \\
& & \\
& = & \dsp{
\frac{1}{4m^2}\, \lh -\Box + m^2 \rh^2\, -\, \frac{\pl^2}{\pl e^2}. }
\ea
\label{5.14}
\ee
\nit
This operator is positive definite, and its zero-modes are zero-modes
of the two terms on the right-hand side separately. Therefore
\be
\Del_{BRST}\, \Psi_{phys}\, =\, 0 \hspace{1em} \Leftrightarrow \hspace{1em}
\left\{ \lh -\Box + m^2 \rh\, \Psi_{phys}\, =\, 0\, \wedge\,
\dd{}{e} \Psi_{phys}\, =\, 0 \right\}.
\label{5.15}
\ee
\nit
Hence the wave functions $\Psi_{phys}$ indeed satisfy the Klein-Gordon
equation and are independend of the gauge variable $e$.
\section{Gauge-fixed propagator and path integral}{\label{S.6}}
The next step is to construct the propagator for the scalar particle
as described by the wave functions (\ref{5.2.2}) in the BRST-extended
state space, using the formalism of sect.(\ref{S.2}). Let us label the
co-ordinates $(x^{\mu},e,b,c)$ collectively by $Z$. Then the
Schr\"{o}dinger equation for the kernel $K_{gf}(Z;Z^{\prime}|T)$ in
the co-ordinate picture becomes:
\be
i \dd{}{T}K_{gf}(Z;Z^{\prime}|T)\, =\, \lh \frac{e}{2m} \lh -\Box +
m^2 - i\ve \rh + i \frac{\pl^2}{\pl b\pl c} \rh\, K_{gf}(Z;Z^{\prime}|T).
\label{6.1}
\ee
\nit
This kernel is required to satisfy the initial condition
\be
\ba{lll}
K_{gf}(Z;Z^{\prime}|0) & = & \del(Z - Z^{\prime}) \\
& & \\
& = & \del^D(x-x^{\prime}) \del(e-e^{\prime}) \del(c-c^{\prime})
\del(b-b^{\prime}).
\ea
\label{6.2}
\ee
\nit
The solution to this equation is
\be
K_{gf}(Z;Z^{\prime}|T)\, =\,
-iT \del(e-e^{\prime})\, \lh \frac{m}{2\pi eT} \rh^{D/2}\,
e^{i \left[ \frac{m}{2eT} \lh x-x^{\prime} \rh^2 - \frac{eT}{2}\lh m - i\ve
\rh
+ \frac{i}{T} \lh b-b^{\prime} \rh \lh c-c^{\prime} \rh \right]}.
\label{6.3}
\ee
\nit
This is the direct counterpart of eq.(\ref{2.10.2}). It is straightforward
to verify that the initial condition (\ref{6.2}) is satisfied, as is the
composition principle:
\be
\int dZ^{\prime}\, K_{gf}(Z;Z^{\prime}|T^{\prime})\,
K_{gf}(Z^{\prime};Z^{\prime\prime}|T^{\prime\prime})\, =\,
K_{gf}(Z;Z^{\prime\prime}|T^{\prime} + T^{\prime\prime}),
\label{6.4}
\ee
\nit
where the integration measure reads
\be
\int dZ\, =\, i \int d^D x de \int db dc.
\label{6.4.1}
\ee
\nit
The propagator in the extended state space then becomes
\be
\Del_{gf}(Z;Z^{\prime})\, =\, \frac{ie}{2m}\, \int_{0}^{\infty} dT\,
K_{gf}(Z;Z^{\prime}|T).
\label{6.5}
\ee
\nit
It is the solution of the inhomogeneous extended Klein-Gordon equation
\be
\lh - \Box + m^2 - i\ve + \frac{2im}{e} \frac{\pl^2}{\pl b\pl c} \rh\,
\Del_{gf}(Z;Z^{\prime})\, =\, \del(Z - Z^{\prime}).
\label{6.6}
\ee
\nit
Substitution of the expression (\ref{6.3}) for the kernel
$K_{gf}(Z,Z^{\prime}|T)$ in eq.(\ref{6.5}) for the generalized propagator
gives a closed expression for the propagator in co-ordinate space which is
conveniently represented in terms of a momentum integral.
Switching to the Fourier representation of the co-ordinate part and
performing the integral over $T$ yields
\be
\Del_{gf}(x,e,b,c;x^{\prime},e^{\prime},b^{\prime},c^{\prime})\, =\,
\frac{2m}{ie}\, \del \lh e - e^{\prime} \rh\, \int \frac{d^D p}{(2\pi)^D}\,
e^{i p \cdot (x - x^{\prime})}\, \lh \frac{ \dsp{
e^{- \frac{ie}{2m} \lh p^2 + m^2 - i \ve \rh \lh b - b^{\prime} \rh
\lh c - c^{\prime} \rh } }}{ \left[ p^2 + m^2 - i \ve \right]^2 } \rh .
\label{6.9}
\ee
\nit
{}From this expression it is straightforward to obtain the Feynman propagator
by integration over the unphysical degrees of freedom:
\be
\ba{lll}
\Del_F(x-x^{\prime}) & = & \dsp{ \int_{\infty}^{\infty} de \int db \int dc\,
\Del_{gf}(x,e,b,c;x^{\prime},e^{\prime},b^{\prime},c^{\prime}) } \\
& & \\
& = & \dsp{ \int \frac{d^Dp}{(2\pi)^D}\,
\frac{ e^{i p \cdot (x - x^{\prime})} }{p^2 + m^2 - i \ve}.}
\ea
\label{6.10}
\ee
\nit
Alternatively, repeated use of the composition principle can be used to
construct a con\-fig\-ura\-tion space path integral representation of the
propagator. Introducing the Fourier representation for the delta function
$\del(e-e^{\prime})$ repeated use of eq.(\ref{6.4}) gives in explicit
notation:
\be
\Del_{gf}(x,e,b,c;x^{\prime},e^{\prime},b^{\prime},c^{\prime})\, =\,
\frac{ie}{2m}\, \int_0^{\infty} dT \int_{\Gam} \left[ \cD x(\tau)
\cD \lb(\tau) \cD e(\tau) \cD c(\tau) \cD b(\tau) \right]\,
e^{ i S_{gf}[x,\lb,e,b,c] },
\label{6.7}
\ee
\nit
where $S_{gf}$ is the classical gauge-fixed action (\ref{3.10}) and the
functional integral is over all paths $\Gam$ in the configuration space
between $(x^{\prime},e^{\prime},b^{\prime},c^{\prime})$ and $(x,e,b,c)$.
{}From the construction a consistent discrete regularization,
specifying the measure to be used, is obtained:
\be
\ba{l}
\dsp{ \int_{\Gam} \left[ \cD x(\tau) \cD \lb(\tau) \cD e(\tau) \cD c(\tau)
\cD b(\tau) \right]\, e^{ i S_{gf}[x,\lb,e,b,c] }\, = }\\
\\
\dsp{ \hspace{2em} =\, \lim_{N \rightarrow \infty}\,
\int \prod_{n=1}^{N} \left[ d^D x_n d \lb_n de_n d c_n d b_n\,
\frac{i\Del T}{2\pi}\, \lh \frac{m}{2\pi i e_n \Del T} \rh^{D/2} \right]
\times } \\
\\
\dsp{ \hspace{2em} e^{ i \sum_{k=0}^N \Del T \left[ \frac{m}{2e_k}
\lh \frac{x^{\mu}_{k+1} - x^{\mu}_k}{\Del T} \rh^2 - \frac{m}{2} e_k
+ \lb_{k+1} \frac{\lh e_{k+1} - e_k \rh}{\Del T} + i \lh \frac{b_{k+1}
- b_k}{\Del T} \rh \lh \frac{c_{k+1} - c_k}{\Del T} \rh \right] }, }
\ea
\label{6.8}
\ee
\nit
where again $\Del T = T/(N+1)$, with $T$ fixed. Note that the measure
contains a factor $\sqrt{m/2\pi i e \Del T}$ for each integral over a
co-ordinate $x^{\mu}$, and a factor $\sqrt{i\Del T}$ for integration over a
ghost $b$ or $c$. Eqs.(\ref{6.7}) and (\ref{6.8}) give a precise meaning to
relation between the path-integral representation of the propagator in the
extended state space and the BRST-invariant gauge-fixed classical action
(\ref{3.10}).
\section{Fermions}{\label{S.7}}
The Dirac-Feynman propagator for a free fermion is the solution of the
inhomogeneous Dirac equation
\be
\lh \gam \cdot \pl_x + m \rh \, S_F(x - y)\, =\, \del^D(x-y),
\label{7.1}
\ee
\nit
with the same causal boundary conditions as for the scalar particle. The
Dirac matrices $\gam^{\mu}$ form a $D$-dimensional Clifford algebra and are
normalized to satisfy the anti-commutation relations
\be
\left\{ \gam^{\mu}, \gam^{\nu} \right\}\, =\, 2\, \eta^{\mu\nu}.
\label{7.2}
\ee
\nit
The standard Fourier integral representation of the Dirac-Feynman propagator is
\be
S_F(x-y)\, =\, \int \frac{d^D p}{(2\pi)^D}\, \frac{ \lh - i \gam \cdot p
+ m \rh } {p^2 + m^2 - i \ve}\, e^{ i p \cdot (x - y)}.
\label{7.3}
\ee
\nit
It is possible to use the Schwinger procedure described in
sect.\ref{S.2} to construct a path-integral representation of this
propagator as a Clifford-algebra valued object. There is however
another method that is often preferred in applications, especially
when interactions are introduced into the model. This method uses
anti-commuting variables to represent the Clifford algebra \ct{Ber,Fad,BDW};
the application to spinning particles has been studied for example in
refs.\ \ct{hht,brink1,brink2}, \ct{BerMar}-\ct{jw4}. The use of this
method here allows a straightforward path-integral representation
of the Feynman propagator for fermions in terms of bosonic co-ordinates
$x^{\mu}$ and a matching set of fermionic (Grassmann-odd) co-ordinates
$\psi^{\mu}$. However, as the details of the procedure depend on the
number of dimensions, we will from now on choose $D = 4$, and limit
ourselves to that physically relevant case. Modifications to treat
the same problem in another number of dimensions are straightforward
to make.
In order to achieve the transition to anti-commuting variables, we define
a representation of the Clifford algebra in terms of two anti-commuting
variables $(\xi^1, \xi^2)$ by
\be
\ba{ll}
\dsp{ \hgm^1 \,=\, \xi^1 + \dd{}{\xi^1} , } &
\dsp{ \hgm^2 \,=\, -i \lh \xi^1 - \dd{}{\xi^1} \rh, } \\
& \\
\dsp{ \hgm^3 \,=\, \xi^2 + \dd{}{\xi^2} , } &
\dsp{ \hgm^0 \,=\, -i \hgm^4\, =\, \xi^2 - \dd{}{\xi^2} . }
\ea
\label{7.3.1}
\ee
\nit
In addition to the $\hgm^{\mu}$ we also need the usual pseudo-scalar element
of the Clifford algebra:
\be
\ba{lll}
\hgm_5 & = & \dsp{ - \frac{i}{4!}\, \ve_{\mu\nu\kg\lb}
\hgm^{\mu} \hgm^{\nu} \hgm^{\kg} \hgm^{\lb} } \\
& & \\
& = & \dsp{ \lh 2 \xi^1 \dd{}{\xi^1} - 1 \rh
\lh 2 \xi^2 \dd{}{\xi^2} - 1 \rh . }
\ea
\label{7.3.3}
\ee
\nit
Note that in contrast to the other $\hgm^{\mu}$, in this realization
$\hgm_5$ is represented by a Grassmann-even (bosonic) operator. Therefore
we refer to this representation as the {\em bosonic} form of the $\hgm_5$.
In later sections we will also encounter a representation of $\hgm_5$
in terms of a Grassmann-odd operator, which is appropriately refered to
as the {\em fermionic} representation.
Now the following algebraic relations are satisfied by these operators:
\be
\left\{ \hgm^{\mu}, \hgm^{\nu} \right\}\, =\, 2 \eta^{\mu\nu} ,
\label{7.3.2}
\ee
\nit
and
\be
\hgm_5^2\, =\, 1 , \hspace{3em}
\left\{ \hgm_5, \hgm^{\mu} \right\}\, =\, 0.
\label{7.3.4}
\ee
\nit
These operators therefore realize the Clifford algebra of the Dirac matrices
including $\gam_5$.
The operators $(\hgm ^{\mu},\hgm _5)$ act on spinors $\Phi(\xi^1,\xi^2)$,
here defined as functions with a Grassmann polynomial structure
\be
\Fg (\xi^1,\xi^2)\, =\, \fg_2\, +\, \xi^1 \fg_3\, -\, \xi^2 \fg_4\, -\,
\xi^1 \xi^2 \fg_1,
\label{7.4}
\ee
\nit
where all co-efficients are functions either of co-ordinates or momentum.
One can define the usual positive-definite inner product for spinors:
\be
\langle \Phi, \Psi \rangle\, =\, \sum_{\ag}\, \fg^*_{\ag} \psi_{\ag}.
\label{7.3.5}
\ee
\nit
This inner product can be written in the Grassmann representation as
\be
\langle \Phi, \Psi \rangle\, =\, \int \prod_{k} \lh d\xi^k d\bar{\xi}^k \rh\,
e^{\bar{\xi} \cdot \xi}\, \Phi^*(\bar{\xi}) \Psi (\xi),
\label{7.3.5.1}
\ee
\nit
where a second set of Grassmann variables $\bar{\xi}^{1,2}$ has been
introduced as argument of the conjugate wave function, and the star
denotes complex conjugation of the c-number co-efficients of $\Phi (\xi)$
plus reversal of the order of the $\xi^k$. It is straightforward to
check that w.r.t.\ this inner product
\be
\langle \Phi, \dd{}{\xi^k} \Psi \rangle\, =\,
\langle \xi^k \Phi, \Psi \rangle.
\label{7.3.5.2}
\ee
\nit
This result implies that in contrast to the other $\hgm^i$ $(i = 1,2,3)$ and
$\hgm^4$, the time-like operator $\hgm^0$ is not real w.r.t.\ $\langle \Phi,
\Psi \rangle $. This is of course to be expected from the lorentzian signature
of space-time. On the other hand, hermiticity can be restored by defining a
(lorentz invariant) indefinite-metric scalar product $\langle \overline{\Phi},
\Psi \rangle $ using the Pauli-conjugate spinorial wave function
\be
\overline{\Fg}(\xi^1,\xi^2)\, =\, \fg_4^*\, +\, \xi^1 \fg_1^*\, -\,
\xi^2 \fg_2^*\, -\, \xi^1 \xi^2 \fg_3^*.
\label{7.4.2}
\ee
\nit
In general the co-efficients of $\Fg$ are independent complex numbers, in
which case they represent the components of a Dirac spinor $\fg_{\ag}$.
Irreducible Weyl spinors can be obtained as eigenfunctions of $\hgm_5$.
As
\be
\hgm_5\, \Fg(\xi^1,\xi^2)\, =\, \Fg(-\xi^1,-\xi^2) ,
\label{7.4.1}
\ee
\nit
it follows that the Grassmann-even components define a Weyl spinor
of positive chirality, whilst the Grassmann-odd components define
a Weyl spinor of negative chirality. In the representation chosen
here the chirality can therefore be identified with the Grassmann
parity of the spinor $\Fg$.
It is also possible to represent Majorana spinors by requiring $C \Fg =
\bar{\Fg}$, where $\bar{\Fg}$ is the Pauli-conjugate spinor
and $C$ is the (anti-hermitean) charge-conjugation operator
\be
C\, =\, \lh \xi^1 - \dd{}{\xi^1} \rh\, \lh \xi^2 - \dd{}{\xi^2} \rh.
\label{7.4.3}
\ee
\nit
The Majorana constraint results in the component relations $\fg_4 =
\fg_1^*$ and $\fg_3 = - \fg_2^*$. In the following we consider Dirac
spinors unless explicitly stated otherwise.
With the above definitions, the Dirac equation can now be transcribed
as follows: let
\be
\Fg^{\prime}(\xi^1,\xi^2)\, =\,
\lh \hgm \cdot p + m \hgm_5 \rh\, \Fg(\xi^1,\xi^2);
\label{7.5}
\ee
\nit
then the components of $\Fg^{\prime}$ are then related to those of $\Fg$ by
\be
\fg^{\prime}_{\ag}\, =\, \left[ \lh - i \gam \cdot p + m \rh \, \gam_5\,
\right]_{\ag\bg}\, \fg_{\bg},
\label{7.6}
\ee
\nit
in the representation of the Dirac matrices defined by
\be
\ba{ll}
\dsp{ \gam_i\, =\, \left( \ba{cc}
0 & -i \sg_i \\
i \sg_i & 0
\ea \right), } & \dsp{
\gam^4\, =\, i \gam^0\, =\,
\left( \ba{cc}
0 & 1 \\
1 & 0
\ea \right), } \\
& \\
\dsp{ \gam_5\, =\, \left( \ba{cc}
1 & 0 \\
0 & -1
\ea \right). }
\ea
\label{7.7}
\ee
\nit
Since the matrix $\gam_5$ is unitary, the equation $\Fg^{\prime} = 0$
is completely equivalent with the free Dirac equation in momentum space,
and we infer that $S_F(x-y)\gam_5$ can be identified with the inverse of
the Dirac operator $(-i \hgm \cdot \pl + m \hgm_5)$ in co-ordinate space.
In the process of translation we have obtained the following correspondence
\be
\hgm^{\mu}\, \mapsto\, -i \gam^{\mu} \gam_5, \hspace{3em}
\hgm_5\, \mapsto \gam_5.
\label{7.8}
\ee
\nit
The next step is to write the inverse of the Dirac operator in the
Grassmann co-ordinate representation. This is achieved by introducing
the ordered symbol for the operator \ct{Ber,Fad,BDW}. A quick way to derive
the necessary results is the following \ct{jw5}. For a single
Grassmann variable $\xi$ the most general form of a differential operator
is
\be
A\, =\, a_0 + a_1 \xi + a_2 \dd{}{\xi} +
\lh a_3 - a_0 \rh \xi \dd{}{\xi}.
\label{7.9}
\ee
\nit
Consider the action of this operator on an arbitrary function
$f(\xi) = f_0 + f_1 \xi$:
\be
A f(\xi)\, =\, a_0 f_0 + a_2 f_1 + \lh a_1 f_0 + a_3 f_1 \rh \xi.
\label{7.10}
\ee
\nit
The operation of $A$ on $f$ can be represented equivalently in terms of an
integral. First observe that
\be
f(\xi)\, =\, \int d\xi^{\prime}\, \del (\xi^{\prime} - \xi) f(\xi^{\prime})\,
=\, \int d\xi^{\prime} d\bar{\xi}\, e^{\bar{\xi} (\xi^{\prime} - \xi)}
f(\xi^{\prime}).
\label{7.11}
\ee
\nit
Here we have introduced the Fourier representation of the anti-commuting
$\del$-function in terms of a conjugate Grassmann variable $\bar{\xi}$.
The ordered symbol $\bar{A}(\xi,\bar{\xi})$ of $A$ is defined by the relation
\be
A f(\xi)\, =\, \int d\xi^{\prime} d\bar{\xi}\, \bar{A}(\xi,\bar{\xi})
e^{\bar{\xi} (\xi^{\prime} - \xi)} f(\xi^{\prime}).
\label{7.12}
\ee
\nit
Our construction shows, that it is obtained by replacing every $\pl/\pl\xi$ in
$A$ by a $\bar{\xi}$:
\be
\bar{A}(\xi,\bar{\xi})\, =\, a_0 + a_1 \xi + a_2 \bar{\xi} +
\lh a_3 - a_0 \rh \xi \bar{\xi}.
\label{7.13}
\ee
\nit
In the following it is useful to have an expression for the symbol of a product
of operators in terms of the product of their symbols. The relation is given by
the equation
\be
\overline{\left[ AB \right]}(\xi,\bar{\xi})\, =\, \int d\xi^{\prime}
d\bar{\xi}^{\prime}\, e^{\lh \bar{\xi}^{\prime} - \bar{\xi} \rh
\lh \xi^{\prime} - \xi \rh }\, \bar{A}(\xi,\bar{\xi}^{\prime})
\bar{B}(\xi^{\prime}, \bar{\xi}).
\label{7.13.1}
\ee
\nit
It is straightforward to generalize this to a product of an arbitrary number
of operators. In that case a symbolic notation for the integration measure is
used, of the form
\be
\int d^n\xi d^n \bar{\xi}\, \equiv\,
\int \prod_{\ag = 1}^n\, d\xi_{\ag} d\bar{\xi}_{\ag}.
\label{7.16.1}
\ee
\nit
Returning to the case of a spinning particle in 4-dimensional space-time we
add two conjugate variables $(\bar{\xi}^1, \bar{\xi}^2)$ and introduce the
symbols for the operators $(\hgm^{\mu}, \hgm_5)$, rescaled by factors
$1/\sqrt{2m}$ for later convenience:
\be
\ba{ll}
\dsp{ \psi^1\, =\, \frac{1}{\sqrt{2m}}\, \lh \xi^1 + \bar{\xi}^1 \rh, } &
\dsp{ \psi^2\, =\, \frac{-i}{\sqrt{2m}}\, \lh \xi^1 - \bar{\xi}^1 \rh, } \\
& \\
\dsp{ \psi^3\, =\, \frac{1}{\sqrt{2m}}\, \lh \xi^2 + \bar{\xi}^2 \rh, } &
\dsp{ \psi^0\, =\, - i \psi^4\,
=\, \frac{1}{\sqrt{2m}}\, \lh \xi^2 - \bar{\xi}^2 \rh. }
\ea
\label{7.14}
\ee
\nit
and
\be
\psi_5\, =\, \frac{1}{\sqrt{2m}}\, e^{2\, \bar{\xi} \cdot \xi}\,
=\, \frac{1}{\sqrt{2m}}\, \lh 2 \xi^1 \bar{\xi}^1 - 1 \rh
\lh 2 \xi^2 \bar{\xi}^2 - 1 \rh .
\label{7.15}
\ee
\nit
Note again the non-standard feature that, whereas the $\psi^{\mu}$ are
Grassmann-odd, the $\psi_5$ introduced here is Grassmann-even (bosonized).
In terms of these quantities, one can rewrite the Dirac equation in momentum
space as
\be
\int d^2\xi^{\prime} d^2\bar{\xi}\, e^{ \bar{\xi} \cdot
\lh \xi^{\prime} - \xi \rh } \, \left[ p \cdot \psi + m \psi_5 \right]
(\xi,\bar{\xi})\, \Fg(\xi^{\prime})\, =\, 0.
\label{7.16}
\ee
\nit
As in our representation the chirality equals the Grassmann parity, it
follows that terms with different Grassmann parity in the wave function
are mixed by a non-zero mass.
To obtain the propagator in momentum space we have to invert\footnote{Of
course, this inverse is to be interpreted in the usual sense of distributions
to deal with singularities.} the Grassmann integral operator in
(\ref{7.16}). As a first step, we observe that the product rule (\ref{7.13.1})
implies the usual identity
\be
\int d^2\xi^{\prime} d^2\bar{\xi}^{\prime}\,
e^{\lh \bar{\xi}^{\prime} - \bar{\xi} \rh \cdot \lh \xi^{\prime} - \xi \rh }\,
\left[ p \cdot \psi + m \psi_5 \right](\xi,\bar{\xi}^{\prime}) \,
\left[ p \cdot \psi + m \psi_5 \right](\xi^{\prime},\bar{\xi})\,
=\, \frac{p^2 + m^2}{2m}.
\label{7.17}
\ee
\nit
The inverse of the Dirac operator in momentum space can then be written
in an extension of the Schwinger representation as \ct{frad}
\be
\ba{lll}
\left[ S_F\, \hgm_5 \right](p_{\mu};\xi,\bar{\xi}) & = & \dsp{
\frac{-i}{\sqrt{2m}}\, \int_0^{\infty} dT \int d\sg\, e^{-\frac{iT}{2m}\,
\lh p^2 + m^2 - i\ve \rh - \sg\, \lh p \cdot \psi + m \psi_5 \rh }
}\\
& & \\
& = & \dsp{ \sqrt{2m}\;
\frac{\left[ p \cdot \psi + m \psi_5 \right](\xi,\bar{\xi})}{
p^2 + m^2 - i\ve }, }
\ea
\label{7.18}
\ee
\nit
where $T$ is the usual Schwinger proper-time parameter, and $\sg$ is a
Grassmann-odd counterpart. Eq.(\ref{7.17}) then implies that
\be
\sqrt{2m}\, \int d^2\xi^{\prime} d^2\bar{\xi}^{\prime}\,
e^{\lh \bar{\xi}^{\prime} - \bar{\xi} \rh \cdot \lh \xi^{\prime} - \xi \rh }\,
\left[ p \cdot \psi + m \psi_5 \right](\xi,\bar{\xi}^{\prime}) \,
\left[ S_F\, \hgm_5 \right] (p_{\mu};\xi^{\prime},\bar{\xi})\, =\, 1.
\label{7.19}
\ee
\nit
In order to get the expression in the co-ordinate representation, we have to
compute the Fourier transform. We also redefine the Grassmann variable,
making it proportional to proper time $T$: $\sg = T \chi$. Now introduce
an integral kernel:
\be
\ba{lll}
K_{\chi}(x - x^{\prime}; \xi, \bar{\xi} |T) & = & \dsp{
\int \frac{d^4p}{(2\pi)^4}\, e^{i p \cdot (x - x^{\prime})}\,
e^{ -\frac{iT}{2m}\, \lh p^2 + m^2 - i\ve \rh -
T \chi \lh p \cdot \psi + m \psi_5 \rh } }\\
& & \\
& = & \dsp{ -i\lh \frac{m}{2\pi T} \rh^{2}\, e^{ \frac{im}{2T}\,
\lh x - x^{\prime} \rh^2 - \frac{iT}{2}\, (m - i\ve)
- m \chi \psi \cdot \lh x - x^{\prime} \rh -
m T \chi \psi_5 }. }
\ea
\label{7.20}
\ee
\nit
Then the following results hold: \nl
$(i)$ The Feynman propagator for a Dirac fermion can be written as
\be
\left[ S_F \hgm_5 \right](x - y; \xi, \bar{\xi})\, =\,
\frac{-i}{\sqrt{2m}}\, \int_0^{\infty} \frac{dT}{T}\, \int d\chi\,
K_{\chi}(x - y; \xi, \bar{\xi} |T).
\label{7.21}
\ee
\nit
$(ii)$ Huygens' composition principle holds in the form
\be
\ba{l}
\dsp{ \int d^4x^{\prime} \int d^2\xi^{\prime} d^2\bar{\xi}^{\prime}\,
e^{\lh \bar{\xi}^{\prime} - \bar{\xi} \rh \cdot \lh \xi^{\prime} - \xi \rh}\,
K_{\chi}(x-x^{\prime}; \xi, \bar{\xi}^{\prime} |T^{\prime})\,
K_{\chi}(x^{\prime}-x^{\prime\prime}; \xi^{\prime}, \bar{\xi}
|T^{\prime\prime})
} \\
\\
\dsp{ \hspace{5em} =\,
K_{\chi}(x-x^{\prime\prime}; \xi, \bar{\xi}; |T^{\prime}+T^{\prime\prime}). }
\ea
\label{7.22}
\ee
\nit
$(iii)$ The integral kernel satisfies the boundary condition
\be
K_{\chi}(x-y; \xi, \bar{\xi} |0)\, =\, \del^4(x-y).
\label{7.23}
\ee
\nit
Apart from giving the explicit expression (\ref{7.18}) for the Dirac-Feynman
propagator, eq.(\ref{7.21}) can be used to construct a path-integral
representation of the propagator by iteration of eq.(\ref{7.22}). Indeed,
repeated use of this equation leads to the result
\be
\ba{ll}
\dsp{ K_{\chi}(x - y; \xi, \bar{\xi} |T) } & =\;\;
\dsp{ \int \prod_{k = 1}^N \left[ d^4 x_k d^2 \xi_k d^2 \bar{\xi}_k \right]
e^{\frac{1}{2} \sum_{j=1}^N \left[ \lh \bar{\xi}_j - \bar{\xi}_{j-1} \rh
\cdot \xi_j - \bar{\xi}_j \cdot \lh \xi_{j+1} - \xi_j \rh \right] } }\\
& \\
& \dsp{ \times \, e^{+ \frac{1}{2} \lh \bar{\xi}_0 - \bar{\xi}_N \rh \cdot
\xi_{N+1} - \frac{1}{2} \bar{\xi}_0 \cdot \lh \xi_1 - \xi_{N+1} \rh }\,
\prod_{s=0}^N\, K_{\chi}(x_{s+1} - x_s; \xi_{s+1}, \bar{\xi}_s | \Del T), }
\ea
\label{7.24}
\ee
\nit
where $x_0 = y$, $x_{N+1} = x$, $\bar{\xi}_0 = \bar{\xi}$, $\xi_{N+1} = \xi$
and $\Del T = T/(N+1)$. Furthermore
\be
\ba{l}
\prod_{s=0}^N\, K_{\chi}(x_{s+1} - x_s; \xi_{s+1}, \bar{\xi}_s | \Del T)\, = \\
\\
\hspace{4em } =\,
\dsp{ \left[ \frac{1}{i} \lh \frac{m}{2\pi \Del T} \rh^2 \right]^{N+1}\,
e^{ \frac{i}{2}\, \sum_{s=0}^N \Del T \left[ m \lh \frac{x_{s+1} - x_s}{
\Del T} \rh^2 + 2i m \chi \psi_s \cdot \lh \frac{x_{s+1} - x_s}{\Del T}
\rh + 2i m \chi \psi_s^5 - m + i\ve \right] } }
\ea
\label{7.25}
\ee
\nit
In agreement with our earlier definitions, the symbols $\psi_k^{\mu}$ here
are a short-hand notation for
\be
\psi_k^1 = \frac{1}{\sqrt{2m}}\, \lh \xi^1_{k+1} + \bar{\xi}^1_k \rh,
\hspace{3em}
\psi_k^2 = \frac{-i}{\sqrt{2m}}\, \lh \xi^1_{k+1} - \bar{\xi}^1_k \rh,
\label{7.26}
\ee
\nit
with similar espressions for $(\psi^3_k, \psi^0_k)$ in terms of $(\xi^2_{k+1},
\bar{\xi}^2_k)$, and the bosonized $\psi^5_k$ given by
\be
\psi_k^5\, =\, \frac{1}{\sqrt{2m}}\, e^{2 \bar{\xi}_k \cdot \xi_{k+1}}.
\label{7.26.1}
\ee
\nit
In the continuum limit ($N \rightarrow \infty, T$ fixed) the free fermion
propagator can now be written as a path integral
\be
\left[ S_F \hgm_5 \right](x - y; \xi, \bar{\xi})\, =\,
\frac{-i}{\sqrt{2m}}\, \int_0^{\infty} \frac{dT}{T}\, \int d\chi\,
\cD x(\tau) \cD \xi(\tau) \cD \bar{\xi}(\tau)\,
e^{i S_{ferm}\left[ x(\tau),\xi(\tau),\bar{\xi}(\tau)\right]},
\label{7.27}
\ee
\nit
where up to boundary terms
\be
\ba{l}
S_{ferm} \left[ x(\tau),\xi(\tau),\bar{\xi}(\tau)\right]\, =\, \\
\\
\dsp{ \hspace{4em}
\int_0^T d\tau\, \left[ \frac{m}{2}\, \dot{x}_{\mu}^2 - \frac{i}{2}\,
\lh \dot{\bar{\xi}} \cdot \xi - \bar{\xi} \cdot \dot{\xi} \rh +
i m \dot{x} \cdot \chi\, \psi(\xi,\bar{\xi}) + i m \chi\,
\psi_5(\xi,\bar{\xi}) - \frac{m}{2} \right]. }
\ea
\label{7.28}
\ee
\nit
Finally it can be cast into a manifestly Lorentz-invariant form\footnote{
There is a subtlety concerning the Lorentz-invariance of the continuum limit,
as it seems to require that the main contribution to the continuum path
integral comes from paths which are smooth in the sense that for $\Del T
\rightarrow 0$ one has $ \| \xi_{k+1} - \xi_k \| = {\cal O}((\Del
T)^{\frac{1}{2}+p})$ with $p > 0$; from the calculation of explicit examples
directly in the continuum limit this seems to be correct. Some arguments for
a consistent continuum path-integral for simple spin systems, related to the
Duistermaat-Heckman theorem, have been advanced in \ct{ercol}.}
by replacing $(\xi^i(\tau), \bar{\xi}^i(\tau))$ by the vector-like variables
$\psi^{\mu}(\tau)$ defined in eq.(\ref{7.14}):
\be
S_{ferm} \left[ x(\tau),\psi^{\mu}(\tau)\right]\, =\, \frac{m}{2}\,
\int_0^T d\tau\, \left[ \dot{x}_{\mu}^2 + i \psi \cdot \dot{\psi}
+ 2 i \dot{x} \cdot \chi\, \psi + 2 i \chi\, \psi_5 - 1 \right].
\label{7.29}
\ee
\nit
This expression resembles closely the one usually encountered in the
literature \ct{brink1,brink2}, \ct{BerMar}-\ct{jw4}, which is based on
the realization of a (gauge-fixed) local world line supersymmetry.
However, in contrast to the standard approach the present construction
uses a Grassmann-{\em even} $\psi_5$, and as a result inclusion of the
mass-term violates explicit supersymmetry at the classical level, even
though it is realized on the operator level in the quantum theory, as
shown by eq.(\ref{7.17}). In fact, for non-vanishing $m$ the
pseudo-classical action $S_{ferm}$ does not even have a well-defined
Grassmann parity; eq.(\ref{7.4.1}) and the discussion following it makes
clear that this is a direct consequence of the non-conservation of
chirality for massive fermions.
It is shown below, that the discrepancy between the result (\ref{7.28})
and the manifestly supersymmetric spinning particle model has its origin in
a doubling of the number of degrees of freedom in the supersymmetric case,
which is regularly overlooked. More precisely, if one preserves supersymmetry
of the classical action by taking $\psi_5$ to be Grassmann-odd, and if the
classical algebra of Poisson-Dirac brackets is mapped in a straightforward
way to the quantum (anti-)commutation relations, then one does not obtain
the representation of the quantum operator $\hat{\psi}_5$ by the 4-dimensional
Dirac matrix $\gam_5 = -i/4!\, \ve_{\mu\nu\kg\lb} \gam^{\mu} \gam^{\nu}
\gam^{\kg} \gam^{\lb}$. Rather, a Grassmann-odd (fermionic) representation of
the operator $\hat{\psi}_5$ is to be used, anti-commuting with the other
$\hat{\psi}^{\mu}$. Then the equivalent matrix representation of this
operator is a $\gam_5$ taken from the Dirac algebra in six dimensions.
This implies a doubling of the number of degrees of freedom of an irreducible
spinor.
On the other hand, for free massless particles manifest supersymmetry is not
violated even in the present theory; this is because chirality is conserved,
and so only terms of even Grassmann parity occur in the action. One therefore
sees that the apparent violation of manifest world line supersymmetry is a
result of mass generation, and has the same dynamical origin.
\section{Worldline supersymmetry}{\label{S.8}}
In this section we turn to the point-particle model with full classical
super-reparametrization invariance on the world line, to compare it
with our treatment of Dirac fermions in sect.(\ref{S.7}). To realize
complete off-shell supersymmetry\footnote{The term {\em off shell}
implies that the supersymmetry algebra is realized without using
dynamical constraints like the equations of motion.} it is necessary
to introduce three different types of super-multiplets (superfields):
\begin{enumerate}
\item{ The gauge multiplet $(e, \chi)$ consisting of the einbein $e$ and its
superpartner $\chi$, also refered to as gravitino; under local worldline
supersymmetry with parameter $\ve$ the multiplet transforms as
\be
\del e\, =\, -2i \ve \chi, \hspace{3em}
\del \chi\, =\, \frac{d\ve}{d\tau}.
\label{8.1}
\ee
}
\item{ Matter multiplets $(x^{\mu}, \psi^{\mu})$ describing the position
and spin co-ordinates of the particle. The transformation rules are
\be
\del x^{\mu}\, =\, -i \ve \psi^{\mu}, \hspace{3em}
\del \psi^{\mu}\, =\, \ve \frac{1}{e}\, \frac{Dx^{\mu}}{D\tau},
\label{8.2}
\ee
\nit
where the super-covariant derivative is constructed with the gravitino
as the connection:
\be
\frac{Dx^{\mu}}{D\tau}\, =\, \frac{dx^{\mu}}{d\tau}\, +\, i \chi \psi^{\mu}.
\label{8.3}
\ee
}
\item{ One or more fermionic multiplets $(\eta, f)$ with Grassmann-odd $\eta$
and even $f$; it is used in the following as an auxiliary multiplet and its
transformation properties under local supersymmetry are
\be
\del \eta\, =\, \ve f, \hspace{3em}
\del f\, =\, -i \ve \frac{1}{e}\, \frac{D\eta}{D\tau}.
\label{8.4}
\ee
\nit
The super-covariant derivative is formed as before:
\be
\frac{D\eta}{D\tau}\, =\, \frac{d\eta}{d\tau}\, -\, \chi f.
\label{8.5}
\ee
}
\end{enumerate}
\nit
It may be checked that in all cases the commutator of two supersymmetry
transformations gives a local translation (reparametrization) with
parameter $a = (2 i/e) \ve_1 \ve_2$, plus a local supersymmetry transformation
with parameter $\ve^{\prime} = -(2 i/e) \ve_1 \ve_2 \chi$.
The minimal free particle action invariant under these one-dimensional local
supersymmetry transformations is
\be
S_{susy} = \frac{m}{2}\, \int d\tau\, \lh
\frac{1}{e}\, \dot{x}_{\mu}^2 + i \psi \cdot \dot{\psi} +
\frac{2i}{e}\, \chi \psi \cdot \dot{x} + i \eta \dot{\eta} +
2i \chi \eta + e f^2 - 2 e f \rh
\label{8.6}
\ee
\nit
Obviously, the bosonic variable $f$ appearing quadratically without derivatives
does not represent a dynamical degree of freedom and may be eliminated using
its algebraic equation of motion $f = 1$ (which is equivalent to completing
the square). Then the classical action becomes
\be
S_{susy}\, =\, \frac{m}{2}\, \int d\tau\, \lh
\frac{1}{e}\, \dot{x}_{\mu}^2 + i \psi \cdot \dot{\psi} +
+ i \eta \dot{\eta} + \frac{2i}{e}\, \chi \psi \cdot \dot{x}
+ 2 i \chi \eta - e \rh.
\label{8.7}
\ee
\nit
In addition local super-reparametrization invariance may be used to fix the
gauge multiplet $(e,\chi)$ to constant values. In particular, if the proper
time $\tau$ is rescaled by $e$, so as to be measured in the same units as the
laboratory time (effectively setting $e = 1$), and the constant value of
$\chi$ is rescaled by the same amount, then comparison with the fermionic
action (\ref{7.28}) shows that $S_{ferm}$ is formally related to $S_{susy}$
by $\dot{\eta} \rightarrow 0$, $\eta \rightarrow \psi_5$. However, this last
relation is frustrated by the mismatch in the Grassmann parity of the
two quantities; moreover, contrary to $f$ the variable $\eta$ is a dynamical
degree of freedom, associated with an additional two-valued parameter labeling
the wave-functions of the particle. As it couples to the gravitino, and
therefore appears in the first-class constraints of the theory, its dynamics
cannot be taken into account by simply equating it to a constant in the action.
At this point we observe however, that one could add more fermionic multiplets
$(\eta_k, f_k)$ whose auxiliary scalars have vanishing classical value:
\be
\Del S\, =\, \frac{m}{2}\, \int d \tau\, \lh i \eta_k \dot{\eta}_k +
e f_k^2 \rh.
\label{8.7.1}
\ee
\nit
Eliminating the $f_k$ by their algebraic Euler-Lagrange equation $f_k = 0$,
there remain only free fermionic degrees of freedom $\eta_k$ which decouple
from the dynamics because they do not interact with the other physical or the
gauge degrees of freedom. In fact these additional fermions define a $d = 1$
topological field theory, in the sense that their action is reparametrization
invariant without involving the metric (represented by the einbein $e$). It
is also easy to see, that they do not contribute to the hamiltonian of the
theory. \vs
\nit
To construct the canonical quantum theory corresponding to the supersymmetric
action $S_{susy}$, we employ the BRST procedure used before. Besides the
reparametrization ghosts $(b,c)$ and the lagrange multiplier $\lb$ we
introduce commuting supersymmetry ghosts $(\ag,\bg)$ and an anti-commuting
multiplier $s$. The nilpotent BRST variations of the full set of variables
read:
\begin{enumerate}
\item{For the gauge multiplet:
\be
\dL e = \frac{d(ce)}{d\tau} + 2 \ag \chi, \hspace{3em}
\dL \chi = \frac{d(c\chi)}{d\tau} + i \dot{\ag}.
\label{8.8}
\ee
}
\item{For the matter multiplets:
\be
\dL x^{\mu} = c \dot{x}^{\mu} + \ag \psi^{\mu}, \hspace{3em}
\dL \psi^{\mu} = c \dot{\psi}^{\mu} + \frac{i\ag}{e}\, \frac{Dx^{\mu}}{D\tau}.
\label{8.9}
\ee
}
\item{For the fermionic multiplet:
\be
\dL \eta = c \dot{\eta} + i \ag f, \hspace{3em}
\dL f = c \dot{f} + \frac{\ag}{e}\, \frac{D\eta}{D\tau}.
\label{8.10}
\ee
}
\item{For the ghost variables:
\be
\ba{ll}
\dsp{ \dL c = c \dot{c} - \frac{i \ag^2}{e}, } &
\dsp{ \dL \ag = c \dot{\ag} + \frac{\ag^2}{e}\, \chi, } \\
& \\
\dsp{ \dL b = - i \lb, } & \dsp{ \dL \bg = s, } \\
& \\
\dsp{ \dL \lb = 0, } & \dsp{ \dL s = 0. }
\ea
\label{8.11}
\ee
}
\end{enumerate}
\nit
In the classical theory defined by $S_{susy}$ we wish to impose the gauge
conditions
\be
\dot{e}\, =\, 0, \hspace{3em} \dot{\chi}\, =\, 0.
\label{8.13}
\ee
\nit
Two write down a relatively simple BRST-invariant gauge-fixed action,
it is convenient to make the redefinitions:
\be
ec\, \rightarrow\, c, \hspace{3em}
\ag - i c \chi\, \rightarrow\, \ag.
\label{8.12}
\ee
\nit
Then the BRST variations are covariantized and simplified; for example
\be
\ba{ll}
\dL \chi = i \dot{\ag}, & \dL c = - i \ag^2, \hspace{3em} \dL \ag = 0, \\
& \\
\dsp{ \dL x^{\mu} = \frac{c}{e}\, \frac{Dx^{\mu}}{D\tau} + \ag \psi^{\mu}, } &
\dsp{ \dL \psi^{\mu} = \frac{c}{e}\, \frac{D\psi^{\mu}}{D\tau} +
\frac{i\ag}{e}\, \frac{Dx^{\mu}}{D\tau}, } \\
& \\
\dsp{ \dL \eta = \frac{c}{e}\, \frac{D\eta}{D\tau} + i \ag f, } &
\dsp{ \dL f = \frac{c}{e}\, \frac{Df}{D\tau} + \frac{\ag}{e}\,
\frac{D\eta}{D\tau}. }
\ea
\label{8.13.1}
\ee
\nit
In terms of these new variables the BRST-invariant gauge-fixed action reads
\be
S_{gf}\, =\, S_{susy}\, +\, \int_0^T d\tau \lh \lb \dot{e} + i s \dot{\chi}
+ i \dot{b} \dot{c} + \dot{\bg} \dot{\ag} + 2 i \ag \dot{b} \chi \rh,
\label{8.14}
\ee
\nit
where we take $S_{susy}$ as in eq.(\ref{8.7}), in which $f = 1$
has been inserted. This is consistent with the BRST variations (\ref{8.13.1})
and their nilpotency, provided the equation of motion $\dot{\eta} = \chi$ is
used.
The canonical momenta for the physical and gauge variables in this theory are
\be
\ba{ll}
\dsp{ p^{\mu} = \frac{m}{e}\, \frac{Dx^{\mu}}{D\tau} , } & \\
& \\
\dsp{ \pi^{\mu} = - \frac{im}{2}\, \psi^{\mu}, } &
\dsp{ \pi_{\eta} = - \frac{im}{2}\, \eta. } \\
& \\
\dsp{ p_e = \lb, } & \dsp{ \pi_{\chi} = - i s, }
\ea
\label{8.15}
\ee
\nit
This is to be supplemented with the ghost momenta
\be
\ba{ll}
\dsp{ \pi_c = - i \dot{b}, } &
\dsp{ \pi_b = i \lh \dot{c} + 2 \ag \chi \rh, } \\
& \\
\dsp{ p_{\ag} = \dot{\bg}, } &
\dsp{ p_{\bg} = \dot{\ag}. }
\ea
\label{8.16}
\ee
\nit
A minor complication is the appearance of the second-class constraints
for the fermionic momenta $(\pi^{\mu}, \pi_{\eta})$. These are resolved
in the standard way by Dirac's procedure, and one ends up with the gauge-fixed,
unconstrained hamiltonian
\be
H_{gf}\, =\, \frac{e}{2m}\, \lh p_{\mu}^2 + m^2 \rh\, -\,
i \chi \lh \psi \cdot p + m \eta - 2 i \ag \pi_c \rh
-\, i \pi_b \pi_c\, +\, p_{\ag} p_{\bg},
\label{8.17}
\ee
\nit
plus the following Dirac-Poisson bracket for functions $(F, G)$ on the
unconstrained phase-space spanned by $(x^{\mu}, p_{\mu}; \psi^{\mu};
\eta; e, p_e; \chi, \pi_{\chi}; c, \pi_c; b, \pi_b; \ag, p_{\ag}; \bg,
p_{\bg})$:
\be
\ba{lll}
\left\{F, G \right\} & = & \dsp{ \dd{F}{x^{\mu}} \dd{G}{p_{\mu}}\, -\,
\dd{F}{p_{\mu}} \dd{G}{x^{\mu}}\, +\, \dd{F}{e} \dd{G}{p_e}\, -\,
\dd{F}{p_e} \dd{G}{e} } \\
& & \\
& & \dsp{
+\, \dd{F}{\ag} \dd{G}{p_{\ag}}\, -\, \dd{F}{p_{\ag}} \dd{G}{\ag}\, +\,
\dd{F}{\bg} \dd{G}{p_{\bg}}\, -\, \dd{F}{p_{\bg}} \dd{G}{\bg}\, } \\
& & \\
& & \dsp{ +\, (-1)^{a_F} \lh \frac{i}{m}\, \dd{F}{\psi^{\mu}}
\dd{G}{\psi_{\mu}} + \frac{i}{m}\, \dd{F}{\eta} \dd{G}{\eta} +
\dd{F}{\chi} \dd{G}{\pi_{\chi}} + \dd{F}{\pi_{\chi}} \dd{G}{\chi} \rh } \\
& & \\
& & \dsp{ +\, (-1)^{a_F} \lh \dd{F}{c} \dd{G}{\pi_c} + \dd{F}{\pi_c} \dd{G}{c}
\, +\, \dd{F}{b} \dd{G}{\pi_b} + \dd{F}{\pi_b} \dd{G}{b} \rh. }
\ea
\label{8.18}
\ee
\nit
The equations of motion then take the canonical form
\be
\frac{dF}{d\tau}\, =\, \left\{ F, H_{gf} \right\}.
\label{8.19}
\ee
\nit
Similarly, the BRST variation of a phase-space function $F$ can now be
obtained from
\be
\dL F\, =\, (-1)^{a_F} \left\{ F, \Og \right\},
\label{8.20}
\ee
\nit
where $\Og$ is the nilpotent, conserved BRST charge
\be
\ba{c}
\dsp{ \Og\, =\, \frac{c}{2m}\, \lh p^2_{\mu} + m^2 \rh\, +\,
\ag \lh p \cdot \psi + m \eta \rh\, - i \pi_b p_e\, +\,
i p_{\bg} \pi_{\chi}\, -\, i \ag^2 \pi_c, } \\
\\
\dsp{ \left\{ \Og, H_{gf} \right\}\, =\, 0, \hspace{3em}
\left\{ \Og, \Og \right\}\, =\, 0. }
\ea
\label{8.21}
\ee
\nit
Like for the scalar particle, the full hamiltonian $H_{gf}$ is BRST-exact:
\be
H_{gf}\, =\, \left\{ - e \pi_c + i \chi p_{\ag}, \Og \right\}.
\label{8.22}
\ee
\nit
In the BRST cohomology the hamiltonian is therefore equivalent to zero and the
physical particle motions are on the mass-shell.
\section{Quantum supersymmetry}{\label{S.9}}
The BRST-invariant classical action for the supersymmetric theory can now
be taken as the starting point for the construction of a canonical
quantum theory for a free point particle. Later this quantum theory is then
used to construct a propagator, and it is shown that it can be expressed in
path-integral form using precisely the classical action $S_{gf}$. Thus the
correspondence between the classical and quantum theory is established
in both directions. From the canonical formulation it will then be entirely
clear, that for $m \neq 0$ this theory describes a degenerate pair of
Dirac fermions, and that this is completely independend of the BRST
quantization procedure; the origin of the doubling of the degrees of freedom
can in fact be traced to the fermionic dynamical variable $\eta$.
The first step in defining the quantum theory is to postulate a set of
(anti) commutation relations in correspondence with the classical Dirac-Poisson
brackets (\ref{8.18}). In addition to the operator (anti) commutators
(\ref{5.1}), we introduce the operator algebra
\be
\ba{rlrl}
\left\{ \hps_{\mu}, \hps_{\nu} \right\} & \dsp{
=\; \frac{1}{m} \eta_{\mu\nu}, } &
\left\{ \hps_6, \hps_6 \right\} & \dsp{ =\; \frac{1}{m}, } \\
& & & \\
\left\{ \hch, \hat{\pi}_{\chi} \right\} & =\; - i, & & \\
& & & \\
\left[ \hag, \hp_{\ag} \right] & =\; i, &
\left[ \hbg, \hp_{\bg} \right] & =\; i.
\ea
\label{9.1}
\ee
\nit
The notation $\hps_6$ (rather than the more usual $\hps_5$) has been introduced
for the operator corresponding to $\eta$, to emphasize the Clifford algebra
structure of the fermion anti-commutators, but also to avoid thinking of this
operator as the product of the other $\hps^{\mu}$ as in the case of the
Dirac fermion in sect.(\ref{S.7}). The above commutation relations can
be satisfied by choosing the following {\em linear} representation of the
operators: the $\hps^{\mu}$ are realized in terms of two Grassmann variables
$(\xi^1, \xi^2)$ and their derivatives, as in (\ref{7.3.1}) by
\be
\hps_{\mu}\, =\, \frac{1}{\sqrt{2m}}\, \hgm_{\mu}(\xi^1,\xi^2).
\label{9.1.1}
\ee
\nit
As before, the operators $\hps_0$ are hermitean only w.r.t.\ to the
physical indefinite-metric inner-product involving Pauli-conjugate spinors.
Furthermore we introduce Grassmann variables $\xi^3$ and $\chi$,
as well as ordinary real variables $(\ag,\bg)$ and define
\be
\ba{llll}
\hps_6 & \dsp{ =\; \frac{1}{\sqrt{2m}}\, \lh \xi^3 + \dd{}{\xi^3} \rh, } &
\hat{\pi}_{\chi} & \dsp{ =\; - i \dd{}{\chi}, } \\
& & & \\
\hp_{\ag} & \dsp{ =\; - i \dd{}{\ag}, } &
\hp_{\bg} & \dsp{ =\; - i \dd{}{\bg}. }
\ea
\label{9.2}
\ee
\nit
Note that we have a fermionic (Grassmann-odd) representation of $\hps_6$.
As a result, having at our disposition the variable $\xi^3$, one can
define at no expense the additional operator
\be
\hps_5\, =\, - \frac{i}{\sqrt{2m}}\, \lh \xi^3 - \dd{}{\xi^3} \rh,
\label{9.2.1}
\ee
\nit
which anti-commutes with the other $\hps_{M}$. This operator does not
appear in the hamiltonian of the theory, which we define in correspondence
with the classical hamiltonian (\ref{8.17}):
\be
\ba{lll}
\hH_{gf} & = & \dsp{ \frac{e}{2m}\, \lh -\Box + m^2 \rh\, -\,
i \chi \lh -i \hps \cdot \pl + m \hps_6 \rh\,
+\, 2 i \ag \chi \dd{}{c}\, +\, i \dd{^2}{b \pl c}\, -\,
\dd{^2}{\ag \pl \bg} } \\
& & \\
& \equiv & \dsp{ e \hH_0\, - \, i \chi \hQ_+\, +\, 2 i \ag \chi \dd{}{c}\, +\,
i \dd{^2}{b \pl c}\, -\, \dd{^2}{\ag \pl \bg}, }
\ea
\label{9.3}
\ee
\nit
and a nilpotent BRST operator
\be
\ba{lll}
\hOg & = & \dsp{ \frac{c}{2m}\, \lh - \Box + m^2 \rh\, +\,
\ag \lh -i \hps \cdot \pl + m \hps_6 \rh\, +\,
i \dd{^2}{e \pl b}\, -\, i \dd{^2}{\bg \pl \chi}\, -\,
\ag^2 \dd{}{c} } \\
& & \\
& = & \dsp{ c \hH_0\, +\, \ag \hQ_+\, +\, i \dd{^2}{e \pl b}\, -\,
i \dd{^2}{\bg \pl \chi}\, -\, \ag^2 \dd{}{c}. }
\ea
\label{9.4}
\ee
\nit
The short-hand notation in the second line of these equations has been
introduced with a view to the supersymmetry algebra
\be
\hQ_+^2\, =\, \hH_0.
\label{9.5}
\ee
\nit
At this point a remark of caution is appropriate: for $m \neq 0$ the operator
$\hH_0$ has zero-modes only in Minkowski space; but for lorentzian metrics
the operator $\hQ_+$ (essentially the Dirac operator) is not hermitean w.r.t.\
the positive-definite inner-product $\Psi^{\dagger} \Psi$ (only w.r.t.\ the
Lorentz-invariant indefinite inner product $\overline{\Psi} \Psi$). Therefore
the zero-modes of $\hH_0$ (the physical states) are not necessarily zero-modes
of $\hQ_+$: one also finds back the negative-energy states associated with the
vanishing of $\hQ_+^{\dagger}$. Moreover, there is a similar relation for the
Lorentz-invariant operator $\hQ_-$ obtained by replacing $m \rightarrow -m$:
\be
\hQ_-^2\, =\, \hH_0, \hspace{3em}
\hQ_-\, =\, \lh - i \hps \cdot \pl - m \hps_6 \rh.
\label{9.6}
\ee
\nit
The resolution of these difficulties lies of course in the full
quantum-field theoretical treatment; it is of interest to see how the
BRST procedure is implemented in this context, and as a first step we
construct in this paper the free propagator, in sect.(\ref{S.10}).
For our present purpose it is however sufficient to note, that the
BRST-cohomology can still be used to characterize the physical states, in
the following way: since the hamiltonian $H_{gf}$ is a BRST-exact operator:
\be
H_{gf}\, =\, \left\{ e \dd{}{c} - i \chi \dd{}{\ag}, \hOg \right\},
\label{9.7}
\ee
\nit
one can pick a state from each equivalence class of solutions of the
BRST condition
\be
\hOg \Psi_{phys}\, =\, 0,
\label{9.8}
\ee
\nit
by requiring the subsidiary condition
\be
\lh e \dd{}{c} - i \chi \dd{}{\ag} \rh\, \Psi_{phys}\, =\, 0,
\label{9.9}
\ee
\nit
for {\em arbitrary} values of $e$ and $\chi$, as suggested by
super-reparametrization invariance. This amounts actually to two
subsidiary conditions:
\be
\dd{}{c}\, \Psi_{phys}\, =\, 0, \hspace{3em}
\dd{}{\ag}\, \Psi_{phys}\, =\, 0.
\label{9.10}
\ee
\nit
One may think of these conditions as defining the ghost vacuum. It now
follows automatically from (\ref{9.7}) that
\be
H_{gf}\, \Psi_{phys}\, =\, 0.
\label{9.11}
\ee
\nit
Combined with the BRST invariance of the physical states expressed by
(\ref{9.8}), this gives the physical states precisely as solutions of
the Dirac (and consequently Klein-Gordon) equation:
\be
\lh - i \hps \cdot \pl + m \hps_6 \rh\, \Psi_{phys}\, =\, 0.
\label{9.12}
\ee
\nit
The wave functions $\Psi_{phys}$ can be decomposed as
\be
\Psi_{phys}(\xi^1,\xi^2,\xi^3)\, =\, \Phi_1(\xi^1,\xi^2)\, +\,
\xi^3 \Phi_2(\xi^1,\xi^2),
\label{9.13}
\ee
\nit
where each of the terms $\Phi_{1,2}(\xi^1,\xi^2)$ is a 4-component spinor of
the type (\ref{7.4}). Therefore the physical states have eight spinor
components rather than four. Working out the action of the Dirac operator on
these wave functions, eq.(\ref{9.12}), it can be written explicitly in matrix
notation, in terms of ordinary four-dimensional Dirac matrices, as
\be
\lh \ba{cc} i \gam_5 \gam \cdot p & m \\
m & - i \gam_5 \gam \cdot p \\
\ea \rh\, \left[ \ba{c} \Fg_1 \\
\Fg_2 \\
\ea \right]\, =\,
\lh \sum_{\mu = 0}^3\, \Gam_{\mu} p^{\mu}\, +\, m \Gam_6\, \rh\, \Psi\, =\, 0,
\label{9.14}
\ee
\nit
where the $\Gam_M$ are an 8-dimensional representation of the Dirac matrices
in 6-dimensional space-time. Thus the supersymmetric spinning particle can
be thought of as a reduction of a 6-dimensional massless spinor to 4
space-time dimensions, by compactification on a circle in the 6th dimension
$(p^6 = m)$ and trivial in $x^5$ $(p^5 = 0)$. It follows, that the states of
this theory represent a degenerate pair of 4-dimensional massive fermions.
\section{Supersymmetric propagator}{\label{S.10}}
In the co-ordinate representation, the quantum states of the supersymmetric
particle in the full extended state space (including the ghost degrees of
freedom) are represented by wave functions depending on the variables
$ Z = (x^{\mu}, \xi^k, e, \chi, c, b, \ag, \bg)$, with $\mu = 0,...,3$ and
$k = 1,2,3$. To obtain the second quantized propagator for this theory, we
first construct the evolution operator associated with the hamiltonian
$\hH_{gf}$, eq.(\ref{9.3}).
It is actually convenient to do this in two steps: first we construct the
expression for the matrix elements of $\hK(T) = e^{-iT\hH_{gf}}$ restricted to
the space of gauge- and ghost degrees of freedom $z = (e, \chi; c, b; \ag,
\bg)$, leaving the `matter' content $(x^{\mu}, \xi^k)$ unspecified; only then
we specify the precise model for the physical degrees of freedom. The
advantage of this procedure is, that the results are easily applied to other
models, for example particles interacting with background fields.
Our starting point is the general expression (\ref{9.3}) for $\hH_{gf}$:
\[
\hH_{gf}\, =\, e \hH_0\, - \, i \chi \hQ_+\, +\, 2 i \ag \chi \dd{}{c}\, +\,
i \dd{^2}{b \pl c}\, -\, \dd{^2}{\ag \pl \bg}, \]
\nit
where $\hH_0 = \hQ_+^2$. For any such $\hH_0$ and $\hQ_+$ not depending on the
gauge- and ghost degrees of freedom the following results hold:
\begin{enumerate}
\item{ The equation
\be
i \dd{}{T}\, \hK(z,\pz|T)\, =\, \hH_{gf }\, \hK(z,\pz|T),
\label{10.1}
\ee
\nit
with the initial condition
\be
\hK (z,\pz|0)\, =\, \del(e - e^{\prime}) \del(\chi - \chi^{\prime})
\del(c - c^{\prime}) \del(b - b^{\prime}) \del(\ag -\ag^{\prime})
\del(\bg - \bg^{\prime})\, \hI,
\label{10.2}
\ee
\nit
has the solution
\be
\ba{ll}
\hK(z,\pz|T)\:\: = & \dsp{ \frac{1}{2\pi}\, \del(e - e^{\prime})
\del(\chi - \chi^{\prime})\, e^{\frac{i}{T}\, (\ag - \ag^{\prime})
(\bg - \bg^{\prime}) - (\ag + \ag^{\prime}) (b - b^{\prime}) \chi
- \frac{1}{T}\, (b - b^{\prime}) (c - c^{\prime})} } \\
& \\
& \dsp{ \times\, e^{- i T \lh e \hH_0 - i \chi \hQ_+ \rh}. }
\ea
\label{10.3}
\ee
}
\item{ This solution satisfies the composition rule
\be
\int d\pz\, \hK(z,\pz|T^{\prime})\,
\hK(\pz,z^{\prime\prime}|T^{\prime\prime})\, =\,
\hK(z,z^{\prime\prime}|T^{\prime} + T^{\prime\prime}).
\label{10.4}
\ee
}
\end{enumerate}
\nit
We observe that the gauge $(e,\chi) = const.$ is manifestly realized
in the expression (\ref{10.3}). Let us now first derive an operator
expression for the propagator in the general case (before specifying
$\hH_0$ and $\hQ_+$):
\be
\hDel_{gf}(z,z^{\prime})\, =\,
\frac{ie}{2m}\, \int_0^{\infty} dT \hK(z,z^{\prime}|T),
\label{10.4.1}
\ee
\nit
which is a solution of the generalized Klein-Gordon-Dirac equation
\be
\hH_{gf}\, \hDel_{gf}(z,z^{\prime})\, =\, \frac{e}{2m}\, \del(z - \pz) \hI.
\label{10.4.2}
\ee
\nit
To get an explicit expression for $\hDel_{gf}$ we use the following
intermediate results:
\be
e^{\frac{i}{T}\, \lh \ag - \ag^{\prime} \rh \lh \bg - \bg^{\prime} \rh }\,
=\, \frac{T}{2\pi}\, \int dp_{\ag} dp_{\bg}\,
e^{i p_{\ag} \lh \ag - \ag^{\prime} \rh + i p_{\bg} \lh \bg - \bg^{\prime} \rh
- i T p_{\ag} p_{\bg} },
\label{10.4.3}
\ee
\nit
and
\be
e^{- \frac{1}{T}\, \lh b - b^{\prime} \rh \lh c - c^{\prime} \rh }\,
=\, \frac{1}{T}\, \int d\pi_b d\pi_c\,
e^{\lh b - b^{\prime} \rh \pi_b + \pi_c \lh c - c^{\prime} \rh
- T \pi_b \pi_c}.
\label{10.4.4}
\ee
\nit
Note that the factors of $T$ in front of these ghost kernels cancel in
the expression for $\hK(T)$, as expected from supersymmetry. It is now
straightforward to obtain the operator expression for the propagator in the
ghost-momentum space:
\be \ba{l}
\dsp{ \hDel_{gf}(z,z^{\prime})\, =\, \frac{e}{8 \pi^2 m}\,
\del(e - e^{\prime}) \del(\chi - \chi^{\prime})\,
e^{- (\ag + \ag^{\prime}) (b - b^{\prime}) \chi}\, \times } \\
\\
\dsp{ \hspace{1em}
\int dp_{\ag} dp_{\bg} \int d\pi_b d\pi_c\, e^{i p_{\ag} (\ag - \ag^{\prime})
+ i p_{\bg} (\bg - \bg^{\prime}) + (b - b^{\prime}) \pi_b +
\pi_c (c - c^{\prime})}\, \left[e \hH_0 - i \chi \hQ_+ + p_{\ag} p_{\bg}
- i \pi_b \pi_c \right]^{-1}. }
\ea
\label{10.4.5}
\ee
\nit
The matrix elements of the inverse operator inside the square brackets
now have to be computed. Again, we first consider the evolution operator.
For the free particle we use the elementary result
\be
\langle x | e^{- iT \lh e \hH_0 - i \chi \hQ_+ \rh} | x^{\prime} \rangle\,
=\, \int \frac{d^4p}{(2\pi)^2}\, e^{i p \cdot \lh x - x^{\prime} \rh}\,
e^{- \frac{ieT}{2m}\, \lh p^2_{\mu} + m^2 \rh - T \chi \lh \hps \cdot p +
m \hps_6 \rh }.
\label{10.5}
\ee
\nit
In addition we replace the fermionic operators $(\hps^{\mu}, \hps_6)$ by
their symbols as in eq.(\ref{7.14}), supplemented by
\be
\eta\, =\, \frac{1}{\sqrt{2m}}\, \lh \xi^3\, +\, \bar{\xi}^3 \rh,
\label{10.6}
\ee
\nit
where we have returned to the original notation for the additional
fermionic degree of freedom. Similarly, for later convenience, we also
introduce the symbol of the presently redundant operator $\hps_5$,
denoting it by $\eta_1$:
\be
\eta_1\, =\, - \frac{i}{\sqrt{2m}}\, \lh \xi^3\, -\, \bar{\xi}^3 \rh.
\label{10.6.1}
\ee
\nit
Carrying out the integration over momentum $p_{\mu}$ then gives the
expression for the matrix element $K(Z,\pZ|T)$:
\be
\ba{ll}
K_{\ve}(Z,\pZ|T)\:\: = & \dsp{ \frac{1}{2\pi}\, \del(e - e^{\prime})
\del(\chi - \chi^{\prime})\, e^{\frac{i}{T}\, (\ag - \ag^{\prime})
(\bg - \bg^{\prime}) - (\ag + \ag^{\prime}) (b - b^{\prime}) \chi
- \frac{1}{T}\, (b - b^{\prime}) (c - c^{\prime})} } \\
& \\
& \dsp{ \times\, \left[ -i \lh \frac{m}{2\pi e T} \rh^2\,
e^{\frac{im}{2eT}\, \lh x - x^{\prime} \rh^2 - \frac{ieT}{2}\,
\lh m - i\ve \rh - \frac{m}{e}\, \chi \psi \cdot \lh x - x^{\prime} \rh
- m T \chi \eta } \right]. }
\ea
\label{10.7}
\ee
\nit
Note that after rescaling $T$ and $\chi$ by a factor $e$ (or equivalently,
taking $e = 1$), the expression in brackets involving the physical
degrees of freedom is almost identical with the expression (\ref{7.20}) for
the kernel $K_{\chi}(T)$ of the single Dirac fermion, except for the
replacement of the Grassmann-even $\psi_5(\xi^i,\bar{\xi}^i)$, $(i=1,2)$,
by the Grassmann-odd $\eta(\xi^3,\bar{\xi}^3)$.
The expression for the propagator $\Del_{gf}(Z,\pZ)$ in the full ghost-extended
state space is similarly obtained by taking the matrix element of
$\hDel(z,z^{\prime})$; this amounts to making the replacement
\be
\ba{l}
\dsp{ \frac{e}{2m}\, \left[e \hH_0 - i \chi \hQ_+ + p_{\ag} p_{\bg}
- i \pi_b \pi_c \right]^{-1}\, =\,
\frac{ie}{2m}\, \int_0^{\infty} dT\, e^{-iT
\left[e \hH_0 - i \chi \hQ_+ + p_{\ag} p_{\bg} - i \pi_b \pi_c \right]} } \\
\\
\hspace{2em} \rightarrow \dsp{
\int \frac{d^4p}{(2\pi)^2}\, \frac{e^{i p \cdot (x - x^{\prime})}}{
p^2 + m^2 - i \ve + \frac{2m}{e}\, \left[ -i \chi (p \cdot \psi + m \eta )
- i \pi_b \pi_c + p_{\ag} p_{\bg} \right]} }
\ea
\label{10.7.1}
\ee
\nit
in the r.h.s.\ of eq.(\ref{10.4.5}), and interpreting $(\psi^{\mu}, \eta)$ as
the above symbols.
The construction of the path-integral formula for the propagator now
repeats the steps for the Dirac fermion, with the difference that instead of
the symbol of the physical part depending on two canonical pairs of Grassmann
variables $(\xi^k,\bar{\xi}^k)$, there are now three such pairs. Then of
course there is also the difference that we have included here additional
ghost degrees of freedom. However, these do not cause any difficulties.
As a result we obtain
\be
\ba{ll}
\dsp{ K_{\ve}(Z,\pZ |T) } & =\;\; \dsp{ \int \left[ \prod_{k = 1}^N dZ_k
\right]\; e^{\frac{1}{2} \sum_{j=1}^N \left[ \lh \bar{\xi}_j -
\bar{\xi}_{j-1} \rh \cdot \xi_j - \bar{\xi}_j \cdot \lh \xi_{j+1} -
\xi_j \rh \right] } }\\
& \\
& \dsp{ \times \, e^{+ \frac{1}{2} \lh \bar{\xi}_0 - \bar{\xi}_N \rh \cdot
\xi_{N+1} - \frac{1}{2} \bar{\xi}_0 \cdot \lh \xi_1 - \xi_{N+1} \rh }\,
\prod_{s=0}^N\, K_{\ve}(Z_{s+1}, Z_s | \Del T), }
\ea
\label{10.8}
\ee
\nit
where the exponent involves 3 types of $(\xi^k,\bar{\xi}^k)$ and as before
$\Del T = T/(N+1)$.
Representing once more the delta-functions by their Fourier decomposition,
and thereby including integrations over lagrange multipliers $(\lb_k,s_k)$ in
the measure $\prod_k dZ_k$, we can use eq.(\ref{10.7}) to evaluate the product
of $K_{\ve}(Z_{s+1},Z_s|\Delta T)$-factors on the right:
\be
\ba{l}
\dsp{ \prod_{s=0}^N\, K_{\ve}(Z_{s+1}, Z_s | \Del T)\, =\, \left[
\frac{1}{i}\, \lh \frac{m}{4\pi^2 e_0 \Del T} \rh^2 \right]^{N+1}\, \times }
\\
\\
\dsp{ \times
e^{ i \sum_{s=0}^N \Del T \left[ \frac{m}{2 e_s}
\lh \frac{x_{s+1} - x_s}{\Del T} \rh^2 - \frac{e_s}{2} \lh m - i\ve \rh +
i \frac{m}{e_s} \chi_{s+1} \psi_s \cdot \lh \frac{x_{s+1} - x_s}{\Del T} \rh
+ i m \chi_{s+1} \eta_s + \lb_{s+1} \lh \frac{e_{s+1} - e_s}{\Del T} \rh
+ i s_{s+1} \lh \frac{\chi_{s+1} - \chi_s}{\Del T} \rh \right]} } \\
\\
\dsp{ \times
e^{i \sum_{s=0}^N \Del T \left[
i \lh \frac{b_{s+1} - b_s}{\Del T} \rh \lh \frac{c_{s+1} - c_s}{\Del T}\rh +
\lh \frac{\bg_{s+1} - \bg_s}{ \Del T}\rh \lh \frac{\ag_{s+1} - \ag_s}{\Del T}
\rh + i \lh \ag_{s+1} + \ag_s \rh \lh \frac{b_{s+1} - b_{s}}{\Del T} \rh
\chi_{s+1} \right] }
}
\ea
\label{10.9}
\ee
\nit
Here the $\psi_k^{\mu}$ are defined as in eq.(\ref{7.26}), and we take
similarly
\be
\eta_k\, =\, \frac{1}{\sqrt{2m}}\, \lh \xi^3_{k+1} + \bar{\xi}^3_k \rh,
\hspace{3em}
\eta_{1k}\, =\, - \frac{i}{\sqrt{2m}}\, \lh \xi^3_{k+1} - \bar{\xi}^3_k \rh .
\label{10.10}
\ee
\nit
Also, in the front factor on the right-hand side of (\ref{10.9}) we have
included a contribution $(e_0)^{2(N+1)}$ in the denominator, instead of
$\prod_{k=0}^N (e_k)^{2}$, by making use of the fact that $e_k$ is constant:
in the integral its value remains the same from time step to time step.
Finally, the exponent of the finite differences in $(\xi_k, \bar{\xi}_k)$
in eq.(\ref{10.8}) can in the continuum limit be rewritten in terms of the
$\psi^{\mu}$, $\eta$ and $\eta_1$ and their derivatives:
\be
\sum_{j = 1}^N\, \left[ \lh \bar{\xi}_j - \bar{\xi}_{j-1} \rh \cdot \xi_j -
\bar{\xi}_j \cdot \lh \xi_{j+1} - \xi_j \rh \right]\: \rightarrow\:
- m \lh \psi \cdot \dot{\psi} + \eta \dot{\eta} + \eta_1 \dot{\eta}_1 \rh.
\label{10.11}
\ee
\nit
As a result we can now construct the propagator of the theory in the
full ghost-extended state space as
\be
\Del_{gf}(Z,Z^{\prime})\, =\, \frac{ie}{2m}\, \int_0^{\infty} dT\,
\int_{\Gam} DZ(\tau)\, e^{i S_{gf}[Z(\tau)]},
\label{10.12}
\ee
\nit
where integration over the lagrange multipliers $(\lb(\tau),s(\tau))$ is to be
included in the measure $\cD Z(\tau)$, and modulo fermionic boundary terms
the action $S_{gf}$ is that of eq.(\ref{8.14}) with the addition of a single
topological fermion of the kind (\ref{8.7.1}). As has been discussed before,
at the classical level this additional fermion is completely harmless, whilst
in the quantum theory it signifies the doubling of the number of components
of the spinor wave-functions, or equivalently the doubling of the number of
degrees of freedom in the propagator.
\section{Conclusions}{\label{S.11}}
In this paper we have studied the propagators of free spin-0 and spin-1/2
particles and connected them to classical relativistic particle-mechanics
through the path-integral formalism. It has been established that starting
from the known field-theoretical expressions is advantageous, as it can
specify the representation of certain operators to be used, something
which is usually not possible from the canonical `Poisson-bracket to
commutator' quantization prescription. In the case of Dirac fermions,
this has been used to argue in favour of the bosonic representation of the
operator $\hps_5$, implying that non-manifestly supersymmetric models are
prefered to avoid doubling of the spectrum of states. Another way to resolve
this problem would be to project out half of the states by additional
constraints. Such an approach, in which a superselection rule is invoked
to restrict the physical matrix elements to half of the spinor degrees
of freedom, has been attempted for example in \ct{bordi}.
Our results can be generalized to the case of particles moving in
certain background fields: scalar, vector (e.g., electro-magnetic) or
gravitational fields can be included in the path-integral expressions
for the propagators. The constructions we have presented have been
designed in such a way that only minimal additional work is needed
to cover these more general cases. An example is the inclusion of scalar
fields, which by Yukawa couplings may give rise to mass generation through
the mechanism of spontaneous symmetry breaking. The contribution of the
scalar interactions to the classical action (\ref{8.6}) is obtained by
replacing the terms for the fermionic multiplet $(\eta, f)$, which for
free particles read $i \eta \dot{\eta} + 2 i \chi \eta + e f^2 - 2 e f$, by
\be
\Del L_{sc}\, =\, i \eta \dot{\eta}\, +\, e f^2\, -\, 2 g e f \vf(x)\, -\,
2 i g e \eta \psi^{\mu} \pl_{\mu} \vf(x)\, +\, 2 i g \chi \eta \vf(x).
\label{11.1}
\ee
\nit
Here $\vf(x)$ is the scalar field, and $g$ is the Yukawa coupling constant.
Eliminating the auxiliary variable $f$ by splitting off a square, this becomes
\be
\Del L_{sc}\, =\, i \eta \dot{\eta}\, -\, eg^2 \vf^2\, -\,
2 i g e \eta \psi^{\mu} \pl_{\mu} \vf\, +\, 2 i g \chi \eta \vf.
\label{11.2}
\ee
\nit
This result, here obtained through multiplet calculus, agrees with the result
derived by dimensional reduction in \ct{mondr}. Taking $\vf(x) = const.\ \neq
0$ returns us to the original action for a free massive particle, and shows
how masses are generated by spontaneous symmetry breaking. It is quite
straightforward to apply these results to the other actions for scalar
and Dirac particles, by either removing all fermionic degrees of freedom,
or by keeping them whilst replacing the fermionic $\eta$ by the bosonized
$\psi_5$.
Interactions with the electro-magnetic or other vector fields can be
introduced, e.g.\ through minimal coupling, whilst gravitational interactions
result from covariantizing the expressions with respect to general co-ordinate
transformations. A quantum treatment of the point particle in curved space has
been presented in \ct{vn1}. A general discussion of the inclusion of
background fields will be presented in a separate paper \ct{jw8}.
\vs
\nit
{\bf Acknowledgement}
\vs
\nit
The research described in this paper is supported in part by the
Human Capital and Mobility program of the European Union through the
network on Constrained Dynamical Systems.
|
\section{Introduction}
The purpose of this paper is to prove the localization theorem for
torus actions in equivariant intersection theory. Using the theorem
we give another proof of the Bott residue formula for Chern numbers of
bundles on smooth complete varieties. In addition, our
techniques allow us to obtain residue formulas for bundles on a
certain class of singular schemes which admit torus actions. This
class is rather special, but it includes some interesting examples
such as complete intersections (cf. \cite{BFQ}) and Schubert varieties.
Let $T$ be a split torus acting on a scheme $X$. The $T$-equivariant
Chow groups of $X$ are a module over $R_T = Sym(\hat{T})$, where
$\hat{T}$ is the character group of $T$. The localization theorem
states that up to $R_T$-torsion, the equivariant Chow groups of the
fixed locus $X^T$ are isomorphic to those of $X$. Such a theorem is a
hallmark of any equivariant theory. The earliest version (for
equivariant cohomology) is due to Borel \cite{Borel}. Subsequently
$K$-theory versions were proved by Segal \cite{Segal} (in topological
$K$-theory), Quart \cite{Quart} (for actions of a cyclic group), and
Thomason \cite{Thomason} (for algebraic $K$-theory \cite{Thomason}).
For equivariant Chow groups, the localization isomorphism is given by
the equivariant pushforward $i_*$ induced by the inclusion of $X^T$ to
$X$. An interesting aspect of this theory is that the push-forward is
naturally defined on the level of cycles, even in the singular case.
The closest topological analogue of this is equivariant Borel-Moore
homology (see \cite{E-G} for a definition), and a similar proof
establishes localization in that theory.
For smooth spaces, the inverse to the equivariant push-forward can be
written explicitly. It was realized independently by several authors
(\cite{I-N}, \cite{A-B}, \cite{B-V}) that for compact spaces, the
formula for the inverse implies the Bott residue formula. In this
paper, we prove the Bott residue formula for actions of split tori on
smooth complete varieties defined over an arbitrary field, also by
computing $(i_*)^{-1}$ explicitly. Bott's residue formula has been
applied recently in enumerative geometry (cf. \cite{E-S}, \cite{K})
and there was interest in a purely intersection-theoretic proof.
Another application of the explicit formula for $(i_*)^{-1}$ is given
in \cite{E-G2}, where we prove (following Lerman \cite{L}) a residue
formula due to Kalkman.
An obvious problem, which should have applications to enumerative
geometry (see e.g. \cite{K}), is to extend the Bott residue formula to
complete singular schemes. Such a formula can be derived when we have
an explicit description of $(i_*)^{-1}[X]_T$, where $[X]_T$ denotes
the equivariant fundamental class of the whole scheme, as follows: Let
$n = \mbox{dim }X$. If $[X]_T = i_*\alpha$ and $p(E)$ is a polynomial
of weighted degree $n$ in Chern classes of equivariant vector bundles
on $X$, then $\mbox{deg }(p(E) \cap [X])$ can be calculated as the
residue of $\pi_*(i^*(p(E)) \cap \beta)$ where $\pi$ is the
equviariant projection from the $X$ (or the fixed locus) to a point.
This approach does not work for equivariant cohomology, because when
$X$ is singular there is no pushforward from $H^*_G(X) \rightarrow
H^*_G(M)$. However, in $K$-theory, where such pushforwards exist,
similar ideas were used by \cite{BFQ} to obtain Lefschetz-Riemann-Roch
formulas for the action of an automorphism of finite order.
The problem of computing $(i_*)^{-1}$ is difficult, but we can do it
in a certain class of singular examples, in particular, if there is an
equivariant embedding $X \stackrel{f}\hookrightarrow M$
into a smooth variety, and every
component of $X^T$ is a component of $M^T$. This condition is
satisfied if $X \subset {\Bbb P}^r$ is an invariant subvariety where $T$
acts linearly with distinct weights (and thus isolated fixed points)
or if $X$ is a Schubert variety in $G/B$. In this context we give a
formula (Proposition \ref{xxxsing}) formula for $(i_*)^{-1}\alpha$ in
terms of $f_*\alpha \in A_*^T(M)$. The case of Schubert
of varieties is worked out in detail in Section \ref{schubs}.
As a consequence it is possible to compute Chern numbers of bundles on
$X$ provided we know $f_*[X]_T \in A_*^T(M)$. Thus for example, if $X$
is a $T$-invariant projective variety and $T$ acts linearly with
distinct weights on ${\Bbb P}^n$, then we can calculate Chern numbers,
provided we know the equivariant fundamental class of $X$. Rather
than write down a general formula, we illustrate this with an
example: in Section \ref{singex} we use a residue calculation to show
that
$$
\int_Q c_1(\pi^* T_{{\Bbb P}^2}) c_1(f^*T_{{\Bbb P}^3}) = 24
$$
where $Q \stackrel{f} \hookrightarrow {\Bbb P}^3$ is a (singular) quadric cone,
and
$\pi: Q \rightarrow {\Bbb P}^2$ is the projection from a point not on $Q$.
The methods can be applied in other examples.
{\bf Acknowledgements.} We thank
Steven Kleiman for suggesting the problem of giving an
algebraic proof of the Bott residue formula. We are also
grateful to Robert Laterveer for discussions of Gillet's
higher Chow groups.
\section{Review of equivariant Chow groups}
In this section we review some of the equivariant intersection
theory developed in \cite{E-G}. The key to
the theory is the definition of equivariant
Chow groups for actions of linear algebraic groups.
All schemes are assumed to be of finite type
defined over a field of arbitrary characteristic.
Let $G$ be a $g$-dimensional group, $X$ an $n$-dimensional scheme and
$V$ a representation of $G$ of dimension $l$. Assume that there is an
open set $U \subset V$ such that a principal bundle quotient $U
\rightarrow U/G$ exist, and that $V-U$ has codimension more than
$n-i$. Thus the group $G$ acts freely on the product $X \times U$. The
group $G$ acts freely on $X \times U$, and if any one of a number of
mild hypotheses is satisfied then a quotient scheme $X_G = (X \times
U)/G$ exists (\cite{E-G}). In particular, if $G$ is special -- for
example, if $G$ is a split torus, the case of interest in this paper
-- a quotient scheme $X_G$ exists.
\begin{defn}
Set $A_i^G(X) = A_{i+l-g}(X_G)$, where $A_*$ is the usual Chow group.
This definition is independent of the choice of $V$ and $U$ as long as
$V-U$ has sufficiently high codimension.
\end{defn}
{\bf Remark:} Because $X \times U \rightarrow X \times^G U$ is
a principal $G$-bundle, cycles on $X \times^G U$ exactly
correspond to $G$-invariant cycles on $X \times U$. Since
we only consider cycles of codimension smaller
than the dimension of $X \times (V-U)$, we may in fact
view these as $G$-invariant cycles on $X \times V$. Thus
every class in $A_i^G(X)$ is represented by a cycle in
$Z_{i+l}(X \times V)^G$, where $Z_*(X \times V)^G$ indicates
the group of cycles generated by invariant subvarieties.
Conversely, any cycle in $Z_{i+l}(X \times V)^G$ determines
an equivariant class in $A_i^G(X)$.
\medskip
The properties of equivariant intersection Chow groups include the following.
(1) Functoriality for equivariant maps:
proper pushforward, flat pullback, l.c.i pullback, etc.
(2) Chern classes of equivariant bundles operate on equivariant
Chow groups.
(3) If $X$ is smooth of dimension $n$, then we denote $A_{n-i}^G(X)$ as
$A^i_G(X)$. In this case there is an intersection
product $A^i_G(X) \times A^j_G(X) \rightarrow A_G^{i+j}(X)$,
so the groups $\oplus_0^\infty A^i_G(X)$ form a graded ring which
we call the equivariant Chow ring. Unlike, the ordinary case
$A^i_G(X)$ can be non-zero for any $i \geq 0$.
(The existence of an intersection product follows from (1),
since the diagonal $X \hookrightarrow X \times X$
is an equivariant regular embedding when $X$ is smooth.)
(4) Of particular use for this paper is the equivariant self-intersection
formula. If $Y \stackrel{i} \hookrightarrow X$ is a regularly embedded
invariant subvariety of codimension $d$,
then
$$i^*i_*(\alpha) = c_d^G(N_YX) \cap \alpha$$ for any
$\alpha \in A_*^G(Y)$.
\subsection{Equivariant higher Chow groups}
Let $Y$ be a scheme. Denote by $A_i(Y,j)$ the higher
Chow groups of Bloch \cite{Bl} (indexed by dimension)
or the groups $CH_{i,i-j}(X)$ defined in \cite[Section 8]{Gillet}.
Both theories agree with ordinary Chow groups when $j=0$,
and both extend the localization short exact sequence for
ordinary Chow groups. However, in the case of Bloch's
Chow groups the localization exact sequence has only
been proved for quasi-projective varieties. The advantage
of his groups is that are naturally defined in terms
of cycles on $X \times \Delta^j$ (where $\Delta^j$ is an algebraic
$j$-simplex) and are rationally isomorphic to higher $K$-theory.
Both these theories can be extended to the equivariant
setting. We define the higher Chow groups $A_i^G(X,j)$
as $A_{i+l-g}(X_G,j)$ for an appropriate mixed space
$X_G$. Because of the quasi-projective hypothesis
in Bloch's work, Bloch's equivariant higher Chow groups
are only defined for (quasi)-projective varieties with
linearized actions. However, Gillet's are defined
for arbitrary schemes with a $G$-action. We
will use two properties of the higher equivariant
theories.\\
(a) If $E \rightarrow X$ an equivariant vector bundle, then
the equivariant Chern classes $c_i^G(E)$ operate
on $A_*^G(X,j)$.\\
(b) If $U \subset X$ is an invariant open set, then there
is a long exact sequence
$$ \ldots \rightarrow A_i^G(U,1) \rightarrow A_i^G(X-U) \rightarrow
A_i^G(X) \rightarrow A_i^G(U) \rightarrow 0.$$
\section{Localization}
In this section we prove the main theorem
of the paper,
the localization theorem for equivariant Chow groups.
For the remainder of the paper, all tori are assumed
to be split, and the coefficients off all Chow groups are rational.
Let $R_T$ denote the $T$-equivariant Chow ring of a point, and let
$\hat{T}$ be the character group of $T$.
\begin{prop} \cite[Lemma 4]{EGT}
$R_T = Sym(\hat{T})\simeq {\Bbb Q}[t_1, \ldots , t_n]$.
where $n$ is the rank of $T$.
$\Box$ \medskip \end{prop}
{\bf Remark.} The identification $R_T = Sym(\hat{T})$
is given explicitly as follows. If $\lambda \in \hat{T}$ is
a character, let $k_\lambda$ be the corresponding 1-dimensional
representation and let $L_\lambda$ denote the line bundle
$U \times^T k_\lambda \rightarrow U/T$. The map
$\hat{T} \rightarrow R^1_T$ given by $\lambda \mapsto c_1(L_{\lambda})$
extends to a ring isomorphism $Sym(\hat{T}) \rightarrow R_T$.
\medskip
\begin{prop}
If $T$ acts trivially on $X$, then $A_*^T(X) = A_*(X) \otimes
R_T$.
\end{prop}
Proof. If the action is trivial then $(U \times X)/T= U/T \times X$.
The spaces $U/T$ can be taken to be products
of projective spaces, so $A_*(U/T \times X) = A_*(X) \otimes A_*(U/T)$.
$\Box$ \medskip
\medskip
If $T \stackrel{f} \rightarrow S$ is a homomorphism of tori, there
is a pullback $\hat{S} \stackrel{f^*} \rightarrow \hat{T}$. This
extends to a ring homomorphism $Sym(\hat{T}) \stackrel{f^*} \rightarrow
Sym(\hat{S})$,
or in other words, a map $f^*: R_S \rightarrow R_T$.
\begin{lemma} \label{t-map}(cf. \cite{A-B})
Suppose there is a $T$-map
$X \stackrel{\phi} \rightarrow S$.
Then $t\cdot A_*^T(X,m)= 0$ for any $t=f^*s$ with $s \in R_S^+$.
\end{lemma}
Proof of Lemma \ref{t-map}.
Since $A^*_S$ is generated in degree 1, we may
assume that $s$ has degree 1. After clearing denominators
we may assume that $s = c_1(L_s)$ for some line bundle
on a space $U/S$. The action of $t=f^*s$ on $A_*(X_T)$ is just
given by $c_1(\pi_T^*f^*L_s)$ where $\pi_T$ is the
map $U \times^T X \rightarrow U/T$.
To prove the lemma we will show that this bundle is trivial.
First note that $L_s = U \times^S k$ for some action of $S$ on
the one-dimensional vector space $k$.
The pullback bundle on $X_T$ is the line bundle
$$U \times^T(X \times k) \rightarrow X_T$$
where $T$ acts on $k$ by the composition of $f:T \rightarrow S$
with the original $S$-action.
Now define a map
$$\Phi: X_T \times k \rightarrow U \times^T(X \times k)$$
by the formula
$$\Phi(e,x,v) = (e,x,\phi(x)\cdot v)$$
(where $\phi(x) \cdot v$ indicates the original $S$ action).
This map is well defined since
\begin{eqnarray*}
\Phi(et,t^{-1}x,v) & = &(et, t^{-1}x, \phi(t^{-1}x) \cdot v)\\
& = & (et,t^{-1}x,t^{-1} \cdot(\phi(x) \cdot v))
\end{eqnarray*}
as required. This map is easily seen to be an isomorphism
with inverse $(e,x,v) \mapsto (e,x,\phi(x)^{-1} \cdot v)$.
$\Box$ \medskip
\begin{prop} \label{fix}
If $T$ acts on $X$
without fixed points, then there exists $r \in R_T^+$ such that
$r \cdot A_*^T(X,m)= 0$. (Recall that $A_*^T(X,m)$ refers to
$T$-equivariant higher Chow groups.)
\end{prop}
\medskip
Before we prove Proposition \ref{fix}, we state and prove
a lemma.
\begin{lemma} \label{porb}
If $X$ is a variety with an action of a torus $T$, then there is
an open $U \subset X$ so that the stabilizer is constant for
all points of $U$.
\end{lemma}
Proof of Lemma \ref{porb}. Let $\tilde X \rightarrow X$ be the normalization
map. This map is $T$-equivariant and is an isomorphism
over an open set. Thus we may assume $X$ is normal. By Sumihiro's
theorem, the $T$ action on $X$ is locally linearizable, so it
suffices to prove the lemma when $X = V$ is a vector space and
the action is diagonal.
If $V = k^n$, then let $U = (k^*)^n$. The $n$-dimensional
torus ${\bf G}_m^n$ acts transitively on $U$ in the obvious way.
This action commutes with the given action of $T$. Thus the stabilizer
at each closed point of $U$ is the same.
$\Box$ \medskip
Proof of Proposition \ref{fix}. Since $A_*^G(X) = A_*^G(X_{red})$
we may assume $X$ is reduced. Working with each component
separately, we may assume $X$ is a variety. Let $X^0 \subset X$
be the ($G$-invariant) locus of smooth points.
By Sumhiro's theorem \cite{Sumihiro}, the action of
a torus on a normal variety is locally
linearizable (i.e. every point has an affine invariant
neighborhood). Using this theorem it is easy to see
that the set
$X(T_1) \subset X^0$ of points with stabilizer
$T_1$ can be given the structure of a locally closed
subscheme of $X$.
By Lemma \ref{porb} there is
some $T_1$ such that $U= X(T_1)$ is open in $X^0$, and thus in $X$.
The torus $T'=T/T_1$ acts without stabilizers,
but the action of $T'$ on $U$ is not a priori proper. However,
by \cite[Proposition 4.10]{Th1}, we can replace $U$
by a sufficiently small open set so that $T'$ acts freely
on $U$ and a principal bundle quotient $U \rightarrow U/T$
exists. Shrinking $U$ further, we can assume that this
bundle is trivial, so there is a $T$ map $U \rightarrow T'$.
Hence, by the lemma, $t \cdot A^T_*(U) = 0$ for any $t \in A_T^*$
which is pulled back
from $A^*_{T'}$.
Let $Z = X -U$. By induction on dimension, we may assume $p \cdot
A^T_*Z = 0$ for some homogeneous polynomial $p \in R_T$. From the
long exact sequence of higher Chow groups,
$$\ldots A_*^T(Z,m) \rightarrow A_*^T(X,m) \rightarrow A_*^T(U,m) \rightarrow
\ldots$$
it follows that $tp$ annihilates $A_*^T(X)$ where
$t$ is the pullback of a homogeneous element of degree $1$
in $R_S$.
$\Box$ \medskip
If $X$ is a scheme with a $T$-action,
we may put a closed subscheme structure
on the locus $X^T$ of points fixed by
$T$ (\cite{Iversen}).
Now $R_T= Sym(\hat{T})$ is a polynomial ring.
Set ${\cal Q}= (R_T^+)^{-1} \cdot R_T$, where $R_T^+$
is the multiplicative system of homogeneous elements of positive degree.
\begin{thm} \label{lcztn}(localization)
The map $i_*:A_*(X^T) \otimes {\cal Q} \rightarrow A_*^T(X) \otimes
{\cal Q}$ is
an isomorphism.
\end{thm}
\medskip
Proof of Theorem \ref{lcztn}.
By Proposition \ref{fix}, $A_*^T(X - X^T,m) \otimes {\cal Q} = 0$.
Thus by the localization exact sequence
$A_*^T(X^T) \otimes {\cal Q} = A_*^T(X) \otimes {\cal Q}$
as desired. $\Box$ \medskip
{\bf Remark.}
The strategy of the
proof is similar to proofs in other theories, see for example
for \cite[Chapter 3.2]{Hsiang}.
\section{Explicit localization for smooth varieties}
The localization theorem in equivariant cohomology has a more explicit
version for smooth varieties because the
fixed locus is regularly embedded.
This yields an integration formula from which
the Bott residue formula is easily deduced (\cite{A-B}, \cite{B-V}).
In this section we prove the analogous results for equivariant Chow
groups of smooth varieties. Because equivariant Chow theory has
formal properties similar to equivariant cohomology, the arguments are
almost the same as in \cite{A-B}. As before we assume that all tori
are split.
Let $F$ be a scheme with a trivial $T$-action.
If $E \rightarrow F$ is a $T$-equivariant vector bundle on
$F$, then $E$ splits canonically into a direct sum of vector subbundles
$\oplus_{\lambda \in \hat{T}} E_{\lambda}$, where $E_{\lambda}$
consists of the subbundle of vectors in $E$ on which $T$ acts by the
character $\lambda$. The equivariant Chern classes of an
eigenbundle $E_{\lambda}$ are given by the following lemma.
\begin{lemma} \label{l.trivchern}
Let $F$ be a scheme with a trivial $T$-action, and let
$E_{\lambda} \rightarrow F$ be a $T$-equivariant vector bundle of rank
$r$ such
that the action of $T$ on each vector in $E_{\lambda}$ is given by the
character $\lambda$. Then for any $i$,
$$
c^T_i(E_{\lambda}) = \sum_{j \leq i}
\left( \begin{array}{c} r-j \\
i-j \end{array}
\right)
c_j(E_{\lambda}) \lambda^{i-j}.
$$
In particular the component of $c_r^T(E_{\lambda})$ in $R^r_T$ is
given by $\lambda^r$. $\Box$ \medskip
\end{lemma}
As noted above, $A^*_T(F) \supset A^*F \otimes R_T$. The lemma
implies that $c^T_i(E)$ lies in the subring $A^*F \otimes R_T$.
Because $A^N F = 0$ for $N > \mbox{dim }F$, elements of $A^i F$, for
$i>0$, are nilpotent elements in the ring $A^*_T(F)$. Hence an
element $\alpha \in A^d F \otimes R_T$ is invertible in $A^*_T(F)$ if
its component in $A^0 F \otimes R^d_T \cong R^d_T$ is nonzero.
For the remainder of this section $X$ will denote a smooth variety
with a $T$ action.
\begin{lemma} \cite{Iversen}
If $X$ is smooth then the fixed locus
$X^T$ is also smooth. $\Box$ \medskip
\end{lemma}
For each component $F$ of the fixed locus $X^T$
the normal bundle $N_FX$ is a $T$-equivariant vector bundle over $F$.
Note that the action of $T$ on $N_FX$ is non-trivial.
\begin{prop}
If $F$ is a component of $X^T$
with codimension $d$ then $c_d^T(N_FX)$ is invertible
in $A^*_T(F) \otimes {\cal Q}$.
\end{prop}
Proof: By (\cite[Proof of Proposition 1.3]{Iversen}),
for each closed point $f \in F$, the tangent space
$T_fF$ is equal to $(T_fX)^T$, so $T$ acts with non-zero weights on the
normal space $N_f = T_fX/T_fF$. Hence the characters $\lambda_i$
occurring in the
eigenbundle decomposition of $N_FX$ are all non-zero. By the
preceding lemma, the component of $c_d^T(N_FX)$ in $R^d_T$ is nonzero.
Hence $c_d^T(N_FX)$ is invertible
in $A^*_T(F) \otimes {\cal Q}$, as desired. $\Box$ \medskip
\medskip
Using this result we can get, for $X$ smooth,
the following more explicit version
of the localization theorem.
\begin{thm} \label{xxx}(Explicit localization)
Let $X$ be a smooth variety with a torus
action.
Let $\alpha \in A_*^T(X) \otimes {\cal Q}$.
Then $$\alpha = \sum_F
i_{F*}\frac{i^*_F\alpha}{c_{d_F}^T(N_FX)},$$ where the sum is over the
components $F$ of $X^T$ and $d_F$ is the codimension of $F$ in $X$.
\end{thm}
Proof: By the surjectivity part of the localization theorem,
we can write $\alpha = \sum_F
i_{F*}(\beta_F)$. Therefore, $i^*_F\alpha = i^*_Fi_{F*}(\beta_F)$
(the other components of $X^T$ do not contribute); by the
self-intersection formula, this is equal to $ c_{d_F}^T(N_FX) \cdot
\beta_F$. Hence $\beta_F = \frac{i^*_F\alpha}{c_{d_F}^T(N_FX)}$ as
desired. $\Box$ \medskip
{\bf Remark.} This formula is valid, using the virtual normal bundle,
even if $X$ is singular, provided that the embedding of the fixed
locus in $X$ is a local complete intersection morphism. Unfortunately,
this condition is difficult to verify. However, if $X$ is cut by a
regular sequence in a smooth variety, and the fixed points are
isolated, then the methods of \cite[Section 3]{BFQ} can be used to
give an explicit localization formula. A similar remark applies to
the Bott residue formula below.
\medskip
If $X$ is complete, then the projection $\pi_X: X \rightarrow pt$
induces push-forward maps $\pi_{X*}: A^T_* X \rightarrow R_T$ and
$\pi_{X*}: A^T_* X \otimes {\cal Q} \rightarrow {\cal Q}$. There
are similar maps with $X$ replaced with any component $F$ of $X^T$.
Applying $\pi_{X*}$ to both sides of the explicit localization
theorem, and noting that $\pi_{X*} i_{F*} = \pi_{F*} $, we deduce
the ``integration formula'' (cf. \cite[Equation (3.8)]{A-B}).
\begin{cor}
(Integration formula) Let $X$ be smooth and complete, and
let $\alpha \in A_*^T(X) \otimes {\cal Q}$. Then
$$\pi_{X*}(\alpha) = \sum_{F \subset X^T} \pi_{F*}\left(
\frac{i^*_F\alpha}{c_{d_F}^T
(N_FX)}\right)$$
as elements of ${\cal Q}$. $\Box$ \medskip
\end{cor}
\medskip
{\bf Remark.} If $\alpha$ is in the image of the natural map $A_*^T(X)
\rightarrow A_*^T(X) \otimes {\cal Q}$ (which need not be injective),
then the equation above holds in the subring $R_T$ of ${\cal Q}$. The
reason is that the left side actually
lies in the subring $R_T$; hence so does the right side. In the
results that follow, we will have expressions of the form $z = \sum
z_j$, where the $z_j$ are degree zero elements of ${\cal Q}$ whose sum
$z$ lies in the subring $R_T$. The pullback map from equivariant to
ordinary Chow groups gives a map $i^*: R_T = A^T_* (pt) \rightarrow {\Bbb Q}
= A_* (pt)$, which identifies the degree 0 part of $R_T$ with ${\Bbb Q}$.
Since $\sum z_j$ is a degree 0 element of $R_T$, it is identified via
$i^*$ with a rational number. Note that $i^*$ cannot be applied to
each $z_j$ separately, but only to their sum. In the integration and
residue formulas below we will identify the degree 0 part of $R_T$
with ${\Bbb Q}$ and suppress the map $i^*$. \medskip
The preceding corollary yields an integration formula for
an element $a$ of the ordinary Chow group $A_0 X$, provided that $a$ is
the pullback of an element $\alpha \in A^T_0 X$.
\begin{prop}
Let $a \in A_0 X$, and suppose that $a = i^* \alpha$ for $\alpha \in
A^T_0 X$. Then
$$
\mbox{deg }(a) = \sum_F \pi_{F*}\{\frac{i^*_F\alpha}{c_{d_F}^T
(N_FX)} \}
$$
\end{prop}
Proof: Consider the commutative diagram
$$\begin{array}{ccc}
X & \stackrel{i} \hookrightarrow & X_T\\
\downarrow\scriptsize{\pi_X} & & \downarrow\scriptsize{\pi^T_X}\\
\mbox{pt} & \stackrel{i} \rightarrow & U/T .
\end{array}$$
We have $\pi_{X*}(a) = \pi_{X*} i^*(\alpha) = i^* \pi^T_{X*}(\alpha)$.
Applying the integration formula gives the result. $\Box$ \medskip
\subsection{The Bott residue formula}
Let $E_1, \ldots , E_s$ be a $T$-equivariant vector bundles
on a complete, smooth $n$-dimensional variety $X$.
Let $p(x^1_1, \ldots x^1_s,\ldots , x^n_1, \ldots x^n_s)$
be a polynomial of weighted degree $n$,
where $x^i_j$ has weighted degree $i$.
Let $p(E_1, \ldots , E_s)$ denote the polynomial
in the Chern classes of $E_1, \ldots , E_s$ obtained
setting $x^i_j = c_i(E_g)$.
The integration formula above will allow us to compute
$\mbox{deg }(p(E_1, \ldots , E_s) \cap [X])$
in terms of the
restriction of the $E_i$ to $X^T$.
As a notational shorthand, write $p(E)$
for $p(E_1, \ldots, E_s)$ and $p^T(E)$ for the corresponding
polynomial in the $T$-equivariant Chern classes of $E_1, \ldots
, E_r$.
Notice that
$p(E) \cap [X] = i^* (p^T(E) \cap [X_T])$.
We can therefore apply the
preceding proposition to get the Bott residue formula.
\begin{thm} \label{bott}
(Bott residue formula) Let $E_1, \ldots , E_r$ be a $T$-equivariant
vector bundles a complete, smooth $n$-dimensional
variety. Then
$$
\mbox{deg }(p(E) \cap [X]) = \sum_{F \subset X^T}
\pi_{F*}\left(\frac{p^T(E|_{F}) \cap
[F]_T}{c_{d_F}^T (N_FX)} \right).
$$
$\Box$ \medskip
\end{thm}
{\bf Remark.} Using techniques of
algebraic deRham homology, H\"ubl and Yekutieli \cite{H-Y} proved a
version of the Bott residue formula, in characteristic 0,
for the action of any algebraic vector
field with isolated fixed points.
\medskip
By Lemma \ref{l.trivchern} the
equivariant Chern classes $c^T_i(E_j|_{F})$ and $c_{d_F}^T (N_FX)$ can
be computed in terms of the characters of the
torus occurring in the eigenbundle
decompositions of $E_j|_{F}$ and $N_FX$ and the Chern classes of the
eigenbundles.
The above formula can then be readily converted (cf. \cite{A-B}) to more
familiar forms of the Bott residue formula not involving equivariant
cohomology. We omit the details. If the torus $T$ is
1-dimensional, then degree zero elements of ${\cal Q}$ are rational
numbers, and the right hand side of the formula is just a sum of
rational numbers. This is the form of the Bott residue formula which
is most familiar in practice.
\section{Localization and residue formulas for singular varieties}
\label{singex}
In general, the problem of proving localization and residue formulas
on singular varieties seems interesting and difficult. In this
section we discuss what can be deduced from an equivariant embedding
of a singular scheme $X$ into a smooth $M$. The results are not very
general, but (as we show) they can be applied in some interesting
examples, for example, if $X$ is a complete intersection in $M = {\Bbb P}^n$
and $T$ acts on $M$ with isolated fixed points, or if $X$ is a
Schubert variety in $M = G/B$.
The idea of using an embedding into a smooth variety to extract
localization information is an old one. In the case of the action of
an automorphism of finite order, the localization and Lefschetz
Riemann-Roch formulas of \cite{Quart}, \cite{BFQ} on quasi-projective
varieties are obtained by a calculation on ${\Bbb P}^n$. Moreover, as in our
case, the best formulas on singular varieties are obtained when the
embedding into a smooth variety is well understood.
At least in principle, a localization theorem can be deduced if every
component of $X^T$ is a component of $M^T$. This holds, for example,
if the action of $T$ on $M$ has isolated fixed points; or if $X$ is a
toric (resp. spherical) subvariety of a nonsingular toric
(resp. spherical) variety $M$. In particular, the condition holds if
$X$ is a Schubert variety and $M$ is the flag variety. We have the
following proposition.
\begin{prop} \label{xxxsing}
Let $f: X \rightarrow M$ be an equivariant embedding of $X$ in a nonsingular
variety $M$. Assume that every component of $M^T$ which intersects $X$ is
contained in $X$. If $F$ is a component of $X^T$, write $i_F$ for the
embedding of $F$ in $X$, and $j_F$ for the embedding of $F$ in $M$. Then:
$(1)$ $f_*: A_*^T(X) \otimes {\cal Q} \rightarrow A_*^T(M) \otimes {\cal Q}$
is injective.
$(2)$ Let $\alpha \in A_*^T(X) \otimes {\cal Q}$.
Then $$\alpha = \sum_F
i_{F*}\frac{j^*_F f_* \alpha}{c_{d_F}^T(N_FM)},$$ where the sum is over the
components $F$ of $X^T$ and $d_F$ is the codimension of $F$ in $M$.
\end{prop}
Proof: (1) Since the components of $X^T$ are a subset of
the components of $M^T$,
$\oplus_{F \subset X^T} A_*^T(F)$ is an $R_T$-submodule of $\oplus
_{F \subset M^T} A_*^T(F)$. By the localization theorem,
$$\sum_{F \subset X^T} j_{F*}(A_*^T(F)) \otimes {\cal Q} \simeq A_*^T(X)
\otimes {{\cal Q}}$$
and
$$\sum_{F \subset M^T} i_{F*}(A_*^T(F)) \otimes {\cal Q} \simeq A_*^T(M)
\otimes {{\cal Q}}.$$
Since $i_{F*} = f_* i_{F *}$, the result follows.
(2) By (1) it suffices to prove
that $$f_*(\alpha - \sum_F
j_{F*}\frac{i^*_F f_* \alpha}{c_{d_F}^T(N_FM)}) = 0 \in A_*^T(X) \otimes
{\cal Q}.$$
Since $i_{F*} = f_* j_{F*}$ the theorem follows from the explicit
localization theorem applied to the class $f_* \alpha$
on the smooth variety $M$.
$\Box$ \medskip
To obtain a residue formula that computes
Chern numbers of bundles on $X$, we only need to know an expansion
$[X]_T =\sum_{F\subset X^T} i_{F*}(\beta_F)$,
where $\beta_F \in A_*^T(F)$. In this case we obtain the formula
$$
\mbox{deg }(p(E) \cap [X]) = \sum_{F \subset X^T}
\pi_{F*}\left(\frac{p^T(i_F^*(E)) \cap \beta_F}{c_{d_F}^T(N_FM)}\right).$$
In the setting of Proposition \ref{xxxsing}, the classes $\beta_F$ are
given by $i_F^* f_* [X]_T$. To obtain a useful residue formula, we
need to make this expression more explicit. This is most easily done
if we can express $f_* [X]_T$ in terms of Chern classes of naturally
occuring equivariant bundles on $M$. The reason is that the pullback
$i_F^*$ of such Chern classes is often easy to compute, particularly
if $F$ is an isolated fixed point (cf. Lemma. \ref {l.trivchern}).
Indeed, this is why the Bott residue formula is a good calculational
tool in the non-singular case.
Although the conditions to obtain localization and residue formulas
are rather strong, they are satisfied in some interesting cases. We
will consider in detail two examples: complete intersections in
projective spaces, and Schubert varieties in $G/B$. For complete
intersections some intrinsic formulas can be deduced using the virtual
normal bundle (see the remark after Theorem \ref{xxx}). In this
section our point of view for complete intersections is different. We
do not use the virtual normal bundle, but instead use the fact that if
$X \stackrel{f} \hookrightarrow M$ is a complete intersection, it is
easy to calculate $f_*[X]_T \in A_*^T(M)$. As an example of our methods
we do a localization and residue calculation on a singular quadric in
${\Bbb P}^3$.
As a final remark, note that to compute Chern numbers of bundles on
$X$ which are pulled back from $M$, it suffices to know $f_*[X]_T$,
for then we can apply residue formulas on $M$. Information about the
fixed locus in $X$ is irrelevant. The interesting case is when the
bundles are not pulled back from $M$; see the example of the singular
quadric below.
\subsection{Complete intersections in projective space}
For simplicity we consider the case where the dimension of $T$ is $1$.
If $T$ acts on a vector space $V$ with weights $a_0, \ldots, a_n$ then
$A^*_T({\Bbb P}(V)) = {\Bbb Z}[h,t] / \prod (h + a_i t)$. We are interested in
complete intersections $X$ in ${\Bbb P}(V)$ where the functions $f_i$
defining $X$ are, up to scalars, preserved by the $T$-action, i.e.,
$t \cdot f_i = t^{a_i} f_i$. In this case we say $f_i$ has weight $a_i$.
The following lemma is immediate.
\begin{lemma}
Suppose $X$ is a hypersurface in ${\Bbb P}(V)$ defined by a homogeneous
polynomial $f$ of degree $d$ and weight
$a$. Then $[X]_T = d h + a t \in A^*_T({\Bbb P}(V))$. Hence if $X$ is a
complete intersection in ${\Bbb P}(V)$ defined by homogeneous polynomials
$f_i$ of degree $d_i$ and weight $a_i$,
then $[X]_T = \prod (d_i h + a_i t)$. $\Box$ \medskip
\end{lemma}
If $T$ acts on $V$ with distinct weights, then $T$ has isolated fixed
points on $M = {\Bbb P}(V)$, and (trivially) every component of $X_T$ is a
component of $M^T$; so by the preceding discussion there is a useful
residue formula. In particular using a little linear algebra we can
easily obtain a formula for $[X]_T$ in terms of the fixed points in
$X$ and the weights of the action. We omit the details to avoid a
notational quagmire, but the ideas are illustrated in the example
of the singular quadric.
\subsection{Schubert varieties in $G/B$} \label{schubs}
In this section, we work over an algebraically closed field. For
simplicity, we take Chow groups to have rational coefficients, and
let $R = R_T \otimes {\Bbb Q}$ denote the rational equivariant Chow ring
of a point.
Let $G$ be a reductive group and $B$ a Borel subgroup, and ${\cal B} = G/B$
the flag variety. In the discussion below, the smooth variety ${\cal B}$ will
play the role of $M$, and the Schubert variety $X_w$ the role of $X$.
Let $T \subset B$ be a maximal torus. $T$ acts on ${\cal B}$ with finitely
many fixed points, indexed by $w \in W$; denote the corresponding
point by $p_w$. More precisely, if we let $w$ denote both an element
of the Weyl group $W = N(T)/T$ and a representative in $N(T)$, then
$p_w$ is the coset $wB$. The flag variety is a disjoint union of the
$B$-orbits $X_w^0 = B \cdot p_w$. The $B$-orbit $X_w^0$ is called a
Schubert cell and its closure $X_w$ a Schubert variety. If $e$
denotes the identity in $W$ and $w_0$ the longest element of $W$, then
$X_e$ is a point and $X_{w_0} = {\cal B}$.
We have $X_w = \cup_{u \leq w} X_u^0$. The $T$ equivariant Chow group
of $X_w$ is a free $R_T$-module with basis $[X_u]_T$, for $u \leq w$.
Let $j_u: p_u \hookrightarrow {\cal B}$. Fix $w \in W$ and let $f: X_w
\hookrightarrow {\cal B}$. For $u \leq w$ let $i_u: p_u \hookrightarrow X_u$.
If $v \leq w$ let $[X_v]_T$ denote the equivariant fundamental class
of $X_v$ in $A_*^T(X_w)$. We want to make explicit the localization
theorem for the variety $X_w$ (which is singular in general), i.e., to
compute $[X_v]_T$ in terms of classes $i_{u*} \beta_u$.
The (rational) equivariant Chow groups $A^*_T({\cal B})$ can be described as
follows. We consider two maps $\rho_1, \rho_2: R \rightarrow
A^*_T({\cal B})$. The map $\rho_1$ is the usual map $R \rightarrow
A^*_T({\cal B})$ given by equivariant pullback from a point. The definition
of $\rho_2$ is as follows. For each character $\lambda \in \hat{T}$
set $\rho_2(\lambda) = c_1^T(M_\lambda)$ where $M_{\lambda}$ is the
line bundle $G \times^B k_{\lambda} \rightarrow {\cal B}$; extend $\rho_2$
to an algebra map $R \rightarrow A^*_T({\cal B})$. The map $R \otimes_{R^W}
R \rightarrow A^*_T({\cal B})$ taking $r_1 \otimes r_2$ to $\rho_1(r_1)
\rho_2(r_2)$ is an isomorphism (see e.g. \cite{Brion}).
We adopt the convention that the Lie algebra of $B$ contains the
positive root vectors. We can identify $T_{p_w}({\cal B})$ with ${\frak g} /
(\mbox{Ad } w) {\frak b}$. This is a representation of $T$ corresponding to the
$T$-equivariant normal bundle of the fixed point $p_w$. We identify
$A^*_T(p_w) \cong R$. From our description of $T_{p_w}({\cal B})$, we
see that (if $n$ denotes the dimension of ${\cal B}$) $c_n^T(N_{p_w}{\cal B})$ is
the product of the roots in ${\frak g} /(\mbox{Ad } w) {\frak b}$, which is easily
seen to give
$$
c_n^T(N_{p_w}{\cal B}) = c_w := (-1)^n(-1)^w \prod_{\alpha > 0} \alpha.
$$
where $n$ is the number of roots $\alpha >0$.
To obtain a localization formula we also need to know the maps $j_u^* :
A^*_T({\cal B}) \rightarrow A^*_T(p_u)$, where $j_u: p_u \hookrightarrow {\cal B}$
is the inclusion. We have identified $A^*_T({\cal B}) = R \otimes_S R$ and
$A^*_T(p_w) = R$. Thus, we may view $j_u^*$ as a map $R \otimes_S R
\rightarrow R$. There is a natural action of $W \times W$ on $R
\otimes_S R$. Let $m: R \otimes_S R \rightarrow R$ denote the
multiplication map.
\begin{lemma}
For $u \in W$, the map $j_u^*: R \otimes_S R \rightarrow R$ equals the
composition $m \circ (1 \times u)$.
\end{lemma}
Proof: It suffices to show that $j_u^* \rho_1(\lambda) = \lambda$ and
$j_u^* \rho_2(\lambda) = u \lambda$. Now, $j_u^* \rho_1$ is just the
equivariant pullback by the map $p_u \rightarrow pt$. Since this
equivariant pullback is how we identify $A^*_T(p_u) = A^*_T(pt) = R$,
with these identifications, $j_u^*\rho_1$ is the identity map, $j_u^*
\rho_1(\lambda) = \lambda$. Also, by definition $j_u^*
\rho_2(\lambda) = c_1^T(M_{\lambda}|_{p_u})$. As a representation of
$T$, $M_{\lambda}|_{p_u} \cong k_{u \lambda}$, so $c_1^T(M_{\lambda}|_{p_u})
= u \lambda$, as desired. $\Box$ \medskip
If $F \in R \otimes_S R$ is a polynomial set
$F(u) = j_u^*F \in R$.
Recall that we have fixed $w$ and let $f : X_w \rightarrow {\cal B}$ denote
the inclusion; for $v \leq w$, $[X_v]_T$ denotes a class in
$A_*^T(X_w)$. By work of Fulton and Pragacz-Ratajski, for $G$
classical, it is known how to express $f_*[X_v]_T \in A^*_T({\cal B})$ in
terms of the isomorphism $A^*_T({\cal B}) \cong R \otimes_{R^W} R$. More
precisely, Fulton and Pragacz-Ratajski (\cite{Fu1}, \cite{P-R})
define elements in $R
\otimes_{{\Bbb Q}} R$ which project to $f_*[X_v]_T$ in $R \otimes_{R^W} R$.
Let $F_u$ denote either the polynomial defined by Fulton or that
defined by Pragacz-Ratajski. Using these polynomials we can get an
explicit localization formula for Schubert varieties.
\begin{prop}
With notation as above, the class $[X_v]_T$ in $A_*^T(X_w) \otimes
{\cal Q}$ is given by
$$
[X_v]_T = (-1)^n
\frac{1}{\prod_{\alpha > 0} \alpha} \sum_{u \leq v} (-1)^u i_{u*}
\left( F_v(u) \cap [p_u]_T \right).
$$
\end{prop}
Proof: This is an immediate consequence of the preceding discussion and
Proposition \ref{xxxsing}. $\Box$ \medskip
{\bf Remark.} Taking $w = w_0$, so $X_w = {\cal B}$, the above formula is an
explicit inverse to the formula of \cite[Section 6.5, Proposition
(ii)]{Brion}. This shows $\frac{f_w(u)}{\Pi_{\alpha > 0} \alpha}$
is Brion's equivariant multiplicity of $X_w$ at the fixed point $p_u$, and
also links Brion's proposition to \cite[Theorem 1.1]{G}.
\subsection{A singular quadric}
In this section we consider the example of the singular quadric $Q
\stackrel{f} \hookrightarrow {\Bbb P}^3$ defined by
the equation $x_0x_1 + x_2^2 = 0$ (note that we allow the
characteristic to be 2). Let ${\Bbb P}^2 \subset {\Bbb P}^3$ be the hyperplane
defined the equation $x_2 = 0$ and let $\pi: Q \rightarrow {\Bbb P}^2$ be
the projection from $(0,0,1,0)$. As a sample of the kinds of the
residue calculations that are possible, we prove the following
proposition.
\begin{prop}
$$\int_Q c_1(\pi^* T_{{\Bbb P}^2}) c_1(f^*T_{{\Bbb P}^3}) = 24$$.
\end{prop}
Proof. We will prove this by considering the following torus action.
Let $T = {\Bbb G}_m$ act on
${\Bbb P}^3$ with weights $(1,-1,0,a)$, where $a \notin \{0,-1,1\}$. The
quadric is invariant under this action, so $T$ acts on $Q$.
Since $(0,0,1,0)$ is a fixed point, $\pi$ is an equivariant map
where $T$ acts on ${\Bbb P}^2$ with weights $(1,-1,a)$.
Thus $c_1^T(\pi^*T_{{\Bbb P}^2}) c_1^T(f^*T_{{\Bbb P}^3}) \cap [Q]_T$
defines an element of $A_*^T(Q) \otimes {\cal Q}$ which
we will express as a residue in terms of the fixed points
for the action of $T$ on $Q$. To do this
we need to express $[Q]_T$ in terms of the fixed points.
By Proposition \ref{xxxsing} this can be done if we know
$f_*[Q]_T \in A_*^T({\Bbb P}^3)$. Since $Q$ is a quadric
of weighted degree 0 with respect to the $T$-action,
$f_*[Q]_T = 2h \in A_*^T({\Bbb P}^3)$.
Since everything can be done explicitly, we will calculate
more than we need and determine the entire $R_T$-module $A_*^T(Q)$
in terms of the fixed points.
\medskip
{\bf Explicit localization on the singular quadric.}
The quadric has a decomposition into affine cells with one cell in
dimensions 0,1 and 2. The open cell is $Q_0 = \{(1,-x^2,x,y) | (x,y)
\in k^2\}$. In dimension 1 the cell is $l_0 = \{(0,1,0,x) | x \in
k\}$, and in dimension 0, the cell is the singular point
\footnote{In characteristic 2, this point is not an isolated singular
point.} $p_s =
\{(0,0,0,1)\}$. Thus, $A_i(Q) = {\Bbb Z}$ for $i = 0,1,2$ with generators
$[Q]$, $[l]$ and $[p_s]$. Moreover, these cells are $T$-invariant, so
their equivariant fundamental classes form a basis for $A_*^T(Q)$ as
an $R_T = {\Bbb Z}[t]$ module. Let ${\Bbb I}$, $L$, and $P_s$ denote the
corresponding equivariant fundamental classes $[Q]_T,[l]_T$, and
$[p_s]_T$.
There are three fixed points $p_s = (0,0,0,1)$, $p = (1,0,0,0)$ and
$p' = (0,1,0,0)$. These points have equivariant fundamental classes
in $A_*^T(Q^T)$ which we denote by $P_s$, $P$, and $P'$.
By abuse of notation we will not distinguish between
$P_s$ and $i_*(P_s)$.
Both $A_*^T(Q^T)$ and $A_*^T(Q)$ are free $R_T$-modules of rank $3$,
with respective ordered bases $\{P_s,P,P' \}$ and $\{
P_s,L,{\Bbb I} \}$. The map $i_*$ is a linear transformation of these
$R_T$-modules, and we will compute its matrix with respect to these
ordered bases. This matrix can be easily inverted, provided we invert
$t$, and so we obtain $(i_*)^{-1}$.
The equivariant Chow ring of ${\Bbb P}^3$ is given by
$$
A^*_T({\Bbb P}^3) = {\Bbb Z}[t,h]/(h-t)(h+t)h(h+at)
$$
so $A^*_T({\Bbb P}^3)$ is free of rank $4$ over $R_T$, with basis
$\{1,h,h^2,h^3\}$.
To compute $i_*P_s$, $i_*P$, and $i_*P'$, we take advantage of the
fact that the pushforward $f_*: A_*^T(Q) \rightarrow A_*^T({\Bbb P}^3)$ is
injective. Moreover it is straightforward to calculate the
pushforward to ${\Bbb P}^3$ of all the classes in our story. To simplify
the notation, we will use $f_*$ to denote either of the maps $A_*^T(Q)
\rightarrow A_*^T({\Bbb P}^3)$ or $A_*^T(Q^T) \rightarrow A_*^T({\Bbb P}^3)$. We find:
$$\begin{array}{l}
f_*({\Bbb I}) = 2h\\
f_*(L)= (h-t)h\\
f_*(P_s) = h^3 - ht^2\\
f_*(P) = h^3 + (a-1) h^2 t - a h t^2\\
f_*(P') = h^3 + (a+1) h^2 t + a h t^2
\end{array} $$
This implies that
$$\begin{array}{l}
i_*(P_s) = P_s\\
i_*(P) = (a-1)t L + P_s\\
i_*(P') = (a+1)t^2 {\Bbb I} + (a+1)t L + P_s.
\end{array} $$
So the matrix for $i_*^T$ is
$$\left(\begin{array}{ccc} 1 & 1 & 1\\ 0 & (a-1)t & (a+1)t\\ 0 & 0 &
(a + 1)t^2 \end{array} \right).$$
Inverting this matrix we obtain
$$\left(\begin{array}{ccc} 1 & \frac{1}{t(- a)} & \frac{2}{t^2(a^2 -1)} \\
0 & \frac{1}{t(a-1)} & \frac{1}{t^2(1-a)} \\
0 & 0 & \frac{1}{t^2(a+1)} \end{array} \right)
$$
Thus we can write (after supressing the $(i_*)^{-1}$ notation)
$$\begin{array}{l}
P_s = P_s\\
L = \frac{1}{t(a-1)}( -P_s + P)\\
{\Bbb I} =\frac{1}{t^2(a^2-1)}(2P_s - (a+1)P +(a -1)P')
\end{array}. $$
\medskip
{\bf Calculation of Chern numbers}
We now return to the task of computing $c_1(\pi^*T_{{\Bbb P}^2}) c_1(f^*T_{{\Bbb P}^3})
\cap {\Bbb I}$. To simplify notation, set $\alpha_1 := c_1(\pi^*T_{{\Bbb P}^2})$
and $\alpha_2 := c_1(f^*T_{{\Bbb P}^3})$ and $\alpha := \alpha_1 \alpha_2$.
By the calculations above
$$ \alpha_1 \alpha_2 \cap {\Bbb I} = i_*(i^*\alpha_1 i^*\alpha_1 \cap
\frac{t^{-2}}{a^2-1}(2P_s - (a+1)P +(a -1)P')).$$
To compute the class explicitly we must compute the restrictions
of $\alpha_1$ and $\alpha_2$ to each of the fixed points $P_s$, $P$ and
$P'$.
The tangent space to $P_s$ in ${\Bbb P}^3$ has weights $(1-a, -1-a, -a)$.
Thus $\alpha_{2}|_{P_s} = (1-a)t - (1+a)t - at = -3at$.
To compute $\alpha_{1}|_{P_s}$ observe that $P_s$ is the
inverse image of the fixed point $(0,0,1) \in {\Bbb P}^2$. Since
$T_{{\Bbb P}^2}$ has weights $(1-a,-1-a)$ at this point,
$c_1(\pi^*T_{{\Bbb P}_2})|_{P_s} = (1-a)t + (-1 -a)t = -2t$.
The restrictions to the other two fixed points can be calculated
similarly. In particular
$$\begin{array}{ll}
\alpha_{1}|_{P} = (a-3)t \mbox{ }& \alpha_{1}|_{P'} = (a+3)t\\
\alpha_{2}|_{P} = (a- 4)t & \alpha_{2}|_{P'} = (a+4)t
\end{array}
$$
Thus,
$$\alpha \cap {\Bbb I} = \frac{12 a^2}{a^2-1}P_s - \frac{(a-3)(a-4)(a+1)}{a^2-1}P
+ \frac{(a+3)(a+4)(a-1)}{a^2-1}P' \in A_*^T(Q) \otimes {\Bbb Q}.$$
Thus,
$$\begin{array}{ll}
\mbox{deg }(c_1(\pi^*T_{{\Bbb P}^2}) c_1(f^{*}T_{{\Bbb P}^3}) \cap [Q]) & =
\frac{12 a^2}{ a^2 -1 } \;- \;\frac{(a-3)(a-4)(a+1)}{a^2-1} \; + \;
\frac{(a+3)(a+4)(a-1)}{a^2-1}\\
& = 24
\end{array}. $$ $\Box$ \medskip
|
\section*{Figure Captions}
\noindent
\begin{figure}[h]
\caption{
The N\'eel order parameters of $8$,$10$,$16$,$18$,$20$-spin clusters
are plotted versus $N^{-1/2}$ (a) in the bond-random Heisenberg model
for $p=0.0,0.1,0.12,0.15$ (from above to below), (b) in the AF-rich
site-random Heisenberg model for $c=0.0,0.2,0.3,0.35,0.4$ and (c) in
the F-rich site-random Heisenberg model for $c=0.0,0.05,0.1,0.15,0.2$.
The straight lines are drawn by the least-squares fitting of the last
four data of each set.
}
\label{neelfig}
\end{figure}
\noindent
\begin{figure}[h]
\caption{
The SG order parameters of $8$,$10$,$16$,$18$,$20$-spin clusters are
plotted versus $1/N$ in the F-rich site-random Heisenberg model for
$c=0.0,0.2,0.4,0.6$ (from above to below). The straight lines are
drawn by the least-squares fitting of the last four data.
}
\label{sgfig}
\end{figure}
\noindent
\begin{figure}[h]
\caption{
The estimates of the SG order parameter in the thermodynamic limit in
the bond-random model (solid line with crosses), the AF-rich site-random
model (solid line with circles), the F-rich site-random model (solid line
with diamonds) and the asymmetric Mattis model (broken line).
}
\label{sgvalfig}
\end{figure}
\end{document}
|
\section{Introduction}
During the last decade, first-principles total-energy
methods based on density functional theory~(DFT) combined
with the pseudopotential method, have become
established as a major tool in the study of condensed
matter~\cite{dft_pseudo}.
The DFT-pseudopotential approach is now widely used for both static
and dynamic simulations on an enormous range of condensed-matter problems.
However, these methods suffer from a severe drawback in that their
computational cost generally increases as the
cube of the number of atoms in the system. This unfavorable
scaling limits the size of systems that can
be studied with current methods and today's computers to a few
hundred atoms at most. This $O(N^3)$ scaling appears in spite of
the fact that the complexity of the problem increases only
linearly with the system size. This observation suggests
that the unfavorable scaling of current methods is
a consequence of the way in which the electronic structure problem is
being addressed. Conventional methods rely either on diagonalization
of the Hamiltonian or orthonormalization of a set of occupied
orbitals, both of which are intrinsically $O(N^3)$ operations.
It is clear that more efficient methods in which the effort is proportional
to the number of atoms must be possible, and in recent years a
considerable effort has been devoted to finding
such `linear-scaling'
schemes~\cite{yang,baroni:giannozzi,galli:parrinello,mauri:galli:car,mauri:galli,ordejon1,ordejon2,kim:mauri:galli,li:nunes:vanderbilt,qiu,hierse:stechel,hernandez:gillan}.
The earliest linear-scaling scheme appears to be the `divide and
conquer' method of Yang~\cite{yang}. This obtains the electronic
density and hence the total energy by dividing the
system into overlapping sub-systems that can be treated independently.
The density is calculated for each sub-system
with conventional LCAO-DFT. The Hamiltonian for each sub-system, which
includes the potential due to the other sub-systems, is
diagonalized independently, thus avoiding the need to diagonalize
the full Hamiltonian.
This procedure is repeated
until self-consistency is achieved.
Baroni and Giannozzi~\cite{baroni:giannozzi} also proposed a scheme
that directly determines the electron density.
They do this by
discretizing the Kohn-Sham Hamiltonian on a real-space grid, and then
using the recursion method of Haydock, Heine and
Kelly~\cite{haydock:heine:kelly} to obtain the diagonal elements of
the Green's function, from which the electron density can be computed
by contour integration. In this case linear scaling results from the
fact that the continued fraction used to evaluate a particular
diagonal element of the Green's function can be truncated once a certain
neighborhood of each point has
been explored. This neighborhood is independent of the system size
for sufficiently large systems.
More recently, several new schemes that
resemble traditional first-principles methods have been
reported. Galli and Parrinello~\cite{galli:parrinello} pointed out
that some improvement could be achieved in the scaling of a
conventional DFT calculation by requiring spatial localization
of the electronic orbitals.
This localization was achieved by adding certain non-local constraining
terms to the Hamiltonian, or by using a filtering procedure. The total
energy can then be obtained as a functional of the localized orbitals
$| \phi_i \rangle$ and their {\em conjugate\/} orbitals
$\bar{| \phi_i \rangle} = \sum_j S^{-1}_{ji} | \phi_j \rangle$,
but in order to
obtain these conjugate orbitals, the overlap matrix $S$ has to be
inverted. Since spatial localization implies sparsity of $S$,
this can be achieved in $O(N^2)$ operations, so that
some improvement with respect to $O(N^3)$ is obtained. A step
further in this direction was made independently by Mauri, Galli and
Car~\cite{mauri:galli:car,mauri:galli} (hereafter referred to as MGC)
and by Ordej\'{o}n {\em et
al.\/}~\cite{ordejon1,ordejon2}. They introduced a new functional of
the occupied orbitals that possesses the same
ground state as the conventional energy functional, but with the added
advantage of leading naturally to orthogonal orbitals when minimized.
If this new functional is minimized with respect to orbitals that are
constrained to remain localized in chosen regions of space, as
suggested by Galli and Parrinello~\cite{galli:parrinello}, a linear
scaling method results. In the original formulation, the number of
orbitals entering the new functional is
equal to half the number of
electrons in the system. This restriction seems to lead to very slow
convergence, and to the appearance of spurious local minima in the
functional. This problem has been recently overcome by Kim, Mauri and
Galli~\cite{kim:mauri:galli}, by generalizing the functional so that
it depends on an arbitrary number of orbitals.
The linear-scaling scheme most relevant to the present work
is that put forward
by Li, Nunes and
Vanderbilt~\cite{li:nunes:vanderbilt} (hereafter referred
to as LNV) in the context
of tight-binding semi-empirical calculations. In this
method, linear scaling is achieved by taking advantage of the real
space localization properties of the density matrix, $\rho({\bf r},
{\bf r}')$. By introducing a spatial cutoff $R_c$ in $\rho$, such that
$\rho({\bf r}, {\bf r}')$ is set to zero if
$| {\bf r} - {\bf r}' | \geq R_c$,
the number of non-zero elements in $\rho$ increases only linearly with
the system size. The electronic structure problem is then formulated
as a minimization of the total energy with respect to the truncated
density matrix, subject to the constraints of idempotency ($\rho^2 =
\rho$) and correct trace ($2 \, \mbox{Tr} \, \rho = N_e$,
where $N_e$ is the number of electrons). The scheme of LNV
consists of an algorithm for imposing these
constraints that at the same time fulfils the goal of linear scaling.
The idempotency of $\rho$ is the most difficult constraint to impose,
and this scheme achieves it by expressing $\rho$ in terms of an
auxiliary matrix, which we denote in this paper by
$\sigma$. This is subjected to a {\em purifying\/}
transformation due to McWeeny~\cite{mcweeny}. If $\sigma$ is a
near-idempotent matrix, i.e. if its eigenvalues lie close to~0 or~1,
this transformation will return $\rho$ as a more nearly idempotent
matrix, and thus it is possible to minimize the total energy with
respect to $\sigma$ while ensuring the near idempotency of $\rho$.
By construction, the method is
variational (i.e. $\min E(R_c) \geq \min E(\infty)$), and it has been
shown that the convergence of calculated properties with the parameter
$R_c$ is fairly rapid~\cite{li:nunes:vanderbilt,qiu}. It is now being
widely used in tight-binding simulations of large systems.
Recently, the idea of working with the density matrix has been applied
to DFT linear scaling schemes. This has been done
independently by Hierse and Stechel~\cite{hierse:stechel} and by
Hern\'{a}ndez and Gillan~\cite{hernandez:gillan}. In both cases, the
density matrix is represented in real space as:
\begin{eqnarray}
\rho({\bf r},{\bf r}') = \sum_{\alpha \beta} \phi_\alpha ({\bf r}) \,
K_{\alpha \beta} \, \phi_\beta^\ast ({\bf r}') \; ,
\end{eqnarray}
where the $\phi_\alpha$ are a set of localized functions, and
$K_{\alpha \beta}$
is a symmetric matrix. The total energy is expressed in terms
of $\rho ( {\bf r}, {\bf r}^\prime )$, and
minimization is
carried out with respect to both the $\phi_\alpha$ and
the $K_{\alpha \beta}$. Hierse and Stechel~\cite{hierse:stechel}
use a number of functions $\phi_\alpha$ equal to the number of
occupied orbitals, but this restriction is not present in our
scheme. The consequences of this and other differences between the
two methods will be addressed later in this paper.
Previously, only a brief description of our method has
been published~\cite{hernandez:gillan}.
In this paper we give a detailed description of
the method, together with some illustrations of its practical
performance and a discussion of its relation to other methods.
In section 2, the method is outlined and its theoretical
foundations are discussed. The practical implementation of the method
is then described in section 3. The tests we have performed
to probe the practical usefulness of the scheme are presented in
section 4. In section 5, we assess what has been achieved and
we discuss possible future developments,
with particular attention to the problems that need to be overcome
before the method can be generally applied. Some of the
mathematical analysis is reported in an Appendix.
\section{Formulation of DFT in terms of the density matrix}
\label{sec:formulation}
\subsection{Density functional theory}
We need to recall briefly the principles of DFT~\cite{ksh}.
The total energy
$E_{\rm tot}$ of the system of valence electrons and atomic cores is
expressed as:
\begin{equation}
E_{\rm tot} = E_{\rm K} + E_{\rm ps} + E_{\rm H} + E_{\rm xc} +
E_{\rm M} \; ,
\label{eq:totalenergy}
\end{equation}
where the terms on the right are the kinetic, pseudopotential,
Hartree and exchange-correlation energies of the electrons, and
$E_{\rm M}$ is the Madelung energy of the cores. The first two energies
are:
\begin{eqnarray}
E_{\rm K} & = & 2 \sum_{i = 1}^{N} \langle \psi_i |
- \frac{\hbar^2}{2 m} \nabla^2 | \psi_i \rangle \nonumber \\
E_{\rm ps} & = & 2 \sum_{i = 1}^{N} \langle \psi_i |
{\hat{V}}_{\rm ps} | \psi_i \rangle \; ,
\end{eqnarray}
where $\psi_i$ are the Kohn-Sham (KS) orbitals, ${\hat{V}}_{\rm ps}$
is the total pseudopotential operator,
and $N = \frac{1}{2} N_e$ is the number of occupied orbitals.
The energies $E_{\rm H}$ and
$E_{\rm xc}$ can be written in terms of the electron number
density $n ( {\bf r} )$:
\begin{eqnarray}
E_{\rm H} & = & \frac{1}{2} e^2 \int d {\bf r} d {\bf r}^{\prime} \,
n ( {\bf r} ) n ( {\bf r}^{\prime} ) /
| {\bf r} - {\bf r}^{\prime} | \nonumber \\
E_{\rm xc} & = & \int d {\bf r} \, n ( {\bf r} ) \epsilon_{\rm xc}
\left( n ( {\bf r} ) \right ) \; ,
\end{eqnarray}
where for simplicity we assume the local density approximation
(LDA) for $E_{\rm xc}$, with $\epsilon_{\rm xc}$ the exchange-correlation
energy per electron.
The number density is:
\begin{equation}
n ( {\bf r} ) = 2 \sum_{i = 1}^{N} | \psi_i ( {\bf r} ) |^2 \; .
\end{equation}
The important principle for present purposes is that the true
ground-state energy and electron density are obtained by minimizing
$E_{\rm tot}$ with respect to the KS orbitals, subject to the
constraint that the latter are kept orthonormal.
In the standard formulation of DFT, which we have just summarized,
all the occupied orbitals are fully occupied. However, it is frequently
convenient, for physical, computational or formal reasons, to
generalize the theory so that orbitals can be partially occupied.
Spatial orbital $\psi_i ( {\bf r} )$, rather than containing
2 electrons, may now contain $2 f_i$ electrons, where
the occupation number $f_i$ lies in the range $0 \leq f_i \leq 1$.
The number density $n ( {\bf r} )$
now becomes:
\begin{equation}
n ( {\bf r} ) = 2 \sum_{i} f_i | \psi_i ( {\bf r} ) |^2 \; ,
\end{equation}
and the kinetic and pseudopotential energies are:
\begin{eqnarray}
E_{\rm K} & = & 2 \sum_i f_i \langle \psi_i |
- \frac{\hbar^2}{2 m} \nabla^2 | \psi_i \rangle \nonumber \\
E_{\rm ps} & = & 2 \sum_i f_i \langle \psi_i |
{\hat{V}}_{\rm ps} | \psi_i \rangle \; .
\end{eqnarray}
The expressions for $E_{\rm H}$ and $E_{\rm xc}$ in terms of
$n ( {\bf r} )$ are unchanged.
The usual physical reason for making this generalization is
that one wishes to treat the electrons at a non-zero temperature,
in which case the $f_i$ are Fermi-Dirac occupation numbers~\cite{mermin};
computationally, the generalization is sometimes made in order to
get rid of the troublesome discontinuity at the Fermi level
in metallic systems~\cite{gillan:89,grumbach:94}.
Our reason for considering it here is that
it will be relevant to the density matrix formulation.
We shall assume that if $E_{\rm tot}$ is minimized both
with respect to the $\psi_i$ (subject to orthonormality) and
with respect to the $f_i$ (subject to the restriction
$0 \leq f_i \leq 1$ and the condition that the sum $f_i$
be equal to $\frac{1}{2} N_e$),
then we arrive at exactly the ground state
that is obtained by the more usual
minimization with respect to fully occupied states $\psi_i$.
Another way of putting this is that the energy cannot be reduced
below the normal ground state by allowing partial
occupation.
Now we turn to the density matrix, which is defined by
\begin{equation}
\rho ( {\bf r}, {\bf r}^{\prime} ) =
\sum_i f_i \psi_i ( {\bf r} ) \psi_i^{*} ( {\bf r}^{\prime} ) \; .
\label{eq:rho_def}
\end{equation}
It follows from this definition that $\rho ( {\bf r}, {\bf r}^{\prime} )$
is a Hermitian operator whose eigenvalues are all in the interval
$[ 0, 1 ]$. The converse is also true: a Hermitian operator
$\rho ( {\bf r} , {\bf r}^{\prime} )$ whose eigenvalues are $f_i$ and
whose eigenfunctions are $\psi_i ( {\bf r} )$
can be written as in
equation (\ref{eq:rho_def}).
In terms of
such an operator $\rho ( {\bf r} , {\bf r}^{\prime} )$,
let the kinetic energy, pseudopotential energy and number density
be defined as:
\begin{eqnarray}
E_{\rm K} & = & - \frac{\hbar^2}{m} \int d {\bf r} \;
\left( \nabla_r^2
\rho ( {\bf r}, {\bf r}^{\prime} ) \right)_{{\bf r} = {\bf r}^{\prime}}
\nonumber \\
E_{\rm ps} & = & 2 \int d {\bf r} d {\bf r}^{\prime} \;
V_{\rm ps} ( {\bf r}^{\prime}, {\bf r} )
\rho ( {\bf r}, {\bf r}^{\prime} ) \nonumber \\
n ( {\bf r} ) & = & 2 \rho ( {\bf r}, {\bf r} ) \; ,
\label{eq:traces}
\end{eqnarray}
with $E_{\rm H}$ and $E_{\rm xc}$ expressed in the usual way in terms
of $n ( {\bf r} )$. It follows from what we have said before
that if $E_{\rm tot}$ is minimized with respect to
$\rho( {\bf r}, {\bf r}^{\prime} )$ subject to the condition that
the eigenvalues of the latter are in the required interval
and add up to $\frac{1}{2} N_e$, then we arrive at
the usual ground state. This
is the density matrix formulation of DFT.
\subsection{Localization of the density matrix}
Since DFT is variational, any restriction placed on the class of
density matrices $\rho ( {\bf r}, {\bf r}^{\prime} )$ that can
be searched over has the effect of raising the minimum energy
$E_{\rm min}$ above its true ground-state value $E_0$; progressive
relaxation of such a restriction makes $E_{\rm min}$ tend to $E_0$.
Now in general the density matrix in the true ground state tends to
zero as the separation of its arguments $| {\bf r} -
{\bf r}^{\prime} |$ increases. This strongly suggests the usefulness
of estimating $E_0$ by searching over $\rho ( {\bf r}, {\bf r}^{\prime} )$
with the restriction that:
\begin{equation}
\rho ( {\bf r}, {\bf r}^{\prime} ) = 0 \; , \; | {\bf r} -
{\bf r}^{\prime} | > R_c \; ,
\label{eq:dm_cutoff}
\end{equation}
where $R_c$ is a chosen cutoff radius. The resulting estimate
$E_{\rm min} ( R_c )$ will tend to $E_0$ from above as
$R_c \rightarrow \infty$. The manner in which $\rho ( {\bf r},
{\bf r}^{\prime} )$ goes to zero at large separations depends
on the electronic structure of the system, and particularly
on whether there is a gap between the highest occupied and
lowest unoccupied states. It is rigourously established that
in one-dimensional systems having a gap $\rho$ decays
exponentially with separation, while in gapless systems it
decays only as an inverse power~\cite{kohn:59}.
It is presumed that three-dimensional
systems behave similarly. This suggests -- though to our knowledge
it is unproven -- that $E_{\rm min} ( R_c ) \rightarrow E_0$
exponentially for insulators and algebraically for metals.
Clearly in practical calculations we cannot work directly with
a six-dimensional function $\rho ( {\bf r}, {\bf r}^{\prime} )$,
even if it vanishes beyond a chosen radius. It is essential that
$\rho$ be separable, i.e. representable in the form:
\begin{equation}
\rho ( {\bf r}, {\bf r}^{\prime} ) =
\sum_{\alpha \beta} \phi_{\alpha} ( {\bf r} ) K_{\alpha \beta}
\phi_{\beta} ( {\bf r}^{\prime} ) \; .
\label{eq:separable}
\end{equation}
For practical purposes, there must be only a finite number of
$\phi_{\alpha} ( {\bf r} )$ functions, which will be referred
to as support functions. For $\rho$ to be Hermitian, we must
require that the matrix $K_{\alpha \beta}$ be Hermitian.
The restriction to a finite number of support functions is
equivalent to the condition that $\rho$ have only this number
of non-zero eigenvalues, and this is the essence of the separability
requirement. With this, we now have two independent restrictions
on $\rho$: localization and separability.
The localization of $\rho$ can be imposed by requiring that
the support functions be non-zero only within chosen regions, which we
call the support regions, and that the coefficients $K_{\alpha \beta}$
vanish if the separation of the support regions of $\phi_\alpha$
and $\phi_\beta$ exceeds a chosen cutoff.
We now have a general framework for linear-scaling DFT schemes.
In practical calculations, the $\phi_\alpha$ functions will be
represented either as a linear combination of basis functions,
or simply by numerical values on a grid. Either way, the amount of
information contained in a support function will be independent
of the size of the system.
The amount of information in the support
functions will then scale linearly with the size of the system,
and the number of $K_{\alpha \beta}$ coefficients will scale in
the same way. This in turn implies that the electron density
$n( {\bf r} )$ and all the terms in the total energy can be calculated
in a number of operations which scales linearly with system
size.
\subsection{Eigenvalue range of the density matrix}
In this general scheme, the ground state is determined by searching
over support functions and $K_{\alpha \beta}$ matrices. However,
it is essential that this search be confined to those $\phi_\alpha$
and $K_{\alpha \beta}$ for which the eigenvalues of
$\rho ( {\bf r}, {\bf r}^\prime )$ lie in the interval [0,1].
This is a troublesome condition to impose, because we certainly
do not wish to work directly with these eigenvalues. We can achieve
what we want by expressing $\rho$ in a form that satisfies the
condition automatically.
The scheme developed in this paper is the DFT analogue of the tight-binding
scheme of LNV~\cite{li:nunes:vanderbilt}.
We write the density matrix as:
\begin{equation}
\rho = 3 \sigma * \sigma - 2 \sigma * \sigma * \sigma \; ,
\end{equation}
where $\sigma ( {\bf r}, {\bf r}^\prime )$ is an auxiliary function.
(The asterix here indicates the continuum analogue of matrix
multiplication. For arbitrary two-point functions
$A ( {\bf r}, {\bf r}^\prime )$ and $B ( {\bf r}, {\bf r}^\prime )$,
we use the notation $C = A * B$ as a short-hand for the statement:
\begin{equation}
C( {\bf r}, {\bf r}^\prime ) =
\int d {\bf r}^{\prime \prime} \, A ( {\bf r}, {\bf r}^{\prime \prime} )
B ( {\bf r}^{\prime \prime} , {\bf r}^{\prime} ) \; .)
\end{equation}
The reason this works is that the eigenvalues $\lambda_\rho$ of
$\rho$ automatically satisfy $0 \leq \lambda_\rho \leq 1$
provided the eigenvalues $\lambda_\sigma$ of $\sigma$ are in the
range $- \frac{1}{2} \leq \lambda_\sigma \leq \frac{3}{2}$; in addition,
$\lambda_\rho$ has turning points at the values 0 and 1.
Since the ground state is obtained when $\lambda_\rho = 0$ or 1, there
is a natural mechanism whereby variation of $\sigma$ drives $\rho$ towards
idempotency.
To obtain the separable form of $\rho$ (Eq.~(\ref{eq:separable})), we write:
\begin{equation}
\sigma ( {\bf r}, {\bf r}^\prime ) = \sum_{\alpha \beta}
\phi_\alpha ( {\bf r} ) L_{\alpha \beta} \phi_\beta ( {\bf r}^\prime ) \; ,
\label{eq:sigma_def}
\end{equation}
which implies the matrix relation:
\begin{equation}
K = 3LSL - 2LSLSL \; ,
\label{eq:k_of_l}
\end{equation}
where $S_{\alpha \beta}$ is the overlap matrix of support functions:
\begin{equation}
S_{\alpha \beta} = \int d {\bf r} \, \phi_\alpha ( {\bf r} )
\phi_{\beta} ( {\bf r} ) \; .
\end{equation}
The ground state is now obtained by minimizing $E_{\rm tot}$
with respect to the $\phi_\alpha$ and the $L_{\alpha \beta}$
matrix, with the $K_{\alpha \beta}$ matrix given by Eq.~(\ref{eq:k_of_l}).
In the practical calculations reported later, the $\phi_\alpha$ are
non-zero only inside spherical regions of radius $R_{\rm reg}$, and
the $L_{\alpha \beta}$ are non-zero only if the centres of the regions
$\alpha$ and $\beta$ are separated by less than a cutoff distance $R_{\rm L}$.
It will be useful for the purposes of later discussion to note how
a closely related scheme leads back to the
MGC method~\cite{mauri:galli:car}. This
scheme is obtained by writing:
\begin{equation}
\rho = \sigma * ( 2 - \sigma ) \; ,
\end{equation}
where $\sigma$ is required to be positive semi-definite. Since the
eigenvalues $\lambda_\sigma$ can be expressed as $\kappa_\sigma^2$
where $\kappa_\sigma$ is real, the eigenvalues of $\rho$ are given
by:
\begin{equation}
\lambda_\rho = \lambda_\sigma ( 2 - \lambda_\sigma ) =
\kappa_\sigma^2 ( 2 - \kappa_\sigma^2 ) \; .
\end{equation}
This quartic function lies in the range [0,1] for $| \kappa_\sigma |
\leq 2^{1/2}$ and has turning points when $\lambda_\rho = 0$ and 1. This
give an alternative mechanism for driving $\rho$ towards idempotency.
With $\sigma$ given, as before, by Eq.~(\ref{eq:sigma_def}), it is
straightforward to show that $\sigma$ is positive semi-definite
if and only if the matrix $L_{\alpha \beta}$ is positive
semi-definite, and this is equivalent to the condition that
$L_{\alpha \beta}$ be expressible as:
\begin{equation}
L_{\alpha \beta} = \sum_s b_\alpha^{(s)} b_\beta^{(s)} \; .
\end{equation}
The result is that $\sigma ( {\bf r}, {\bf r}^\prime )$ must have
the form:
\begin{equation}
\sigma ( {\bf r}, {\bf r}^\prime ) = \sum_s \chi^{(s)} ( {\bf r} )
\chi^{(s)} ( {\bf r}^\prime ) \; ,
\end{equation}
where:
\begin{equation}
\chi^{(s)} ( {\bf r} ) = \sum_{\alpha} b_\alpha^{(s)}
\phi_\alpha ( {\bf r} ) \; .
\end{equation}
Following arguments presented by Nunes and Vanderbilt~\cite{nunes:vanderbilt}
in the tight-binding
context, it can now be shown that this scheme is exactly
equivalent to the linear-scaling DFT scheme of MGC.
\section{Practical implementation of the method}
\label{sec:details}
\subsection{The real-space grid}
We now give a prescription for the calculation of the energy functional,
and of its derivatives with respect to the support
functions $\phi_\alpha$ and
the $L_{\alpha \beta}$ parameters, and we describe how minimization
of the energy
can be carried out in practice. Central to our implementation of the
method described in the previous section is the use of a regular
cubic real-space grid,
spanning the whole system under study.
There have been a number of recent implementations of conventional
DFT-pseudopotential calculations using real-space
grids~\cite{chelikowsky,briggs,gygi,zumbach}.
The support functions are represented by their values at the grid
points. Since these functions are required to be spatially
localized, they have non-zero values only on the grid points
inside the localization regions. In the present work, these
regions are chosen to be spherical, and their centres are at the atomic
positions. Real-space integration is replaced by summation over grid
points, so that e.g. the overlap matrix elements are calculated as:
\begin{eqnarray}
S_{\alpha \beta} \simeq \delta \omega \sum_{{\bf r}_\ell} \phi_\alpha
({\bf r}_\ell) \, \phi_\beta({\bf r}_\ell) \; ,
\label{eq:overlap}
\end{eqnarray}
where the sum goes over the set of grid points ${\bf r}_\ell$
common to the localization regions of both $\phi_\alpha$
and $\phi_\beta$, and $\delta \omega$ is the volume per grid point.
The action of the kinetic energy operator on the support
functions is evaluated using a finite difference technique.
To $n$th order in the grid spacing, $h$, we have that
\begin{eqnarray}
\frac{\partial^2 \phi_\alpha}{\partial x^2} (n_x, n_y, n_z) \simeq
\frac{1}{h^2} \sum_{m=-n}^{n} C_{\mid m \mid}
\phi_\alpha (n_x+m,n_y,n_z),
\label{eq:finitedifference}
\end{eqnarray}
where $n_x, n_y$ and $n_z$ are integer indices
labelling grid point ${\bf r}_\ell$,
and the coefficients $C_{\mid m \mid}$ can be calculated beforehand.
Equivalent expressions can be used for $\partial^2 \phi_\alpha/\partial y^2$
and $\partial^2 \phi_\alpha /\partial z^2$, and it is thus possible to
evaluate $\nabla^2 \phi_\alpha$ approximately at each grid point.
{}From Eqs.~(\ref{eq:traces}) and (\ref{eq:separable}),
the kinetic energy is given by:
\begin{equation}
E_{\rm K} = 2 \sum_{\alpha \beta} K_{\alpha \beta} T_{\beta \alpha} \; ,
\end{equation}
where
\begin{equation}
T_{\beta \alpha} = - \frac{\hbar^2}{2 m} \int d {\bf r} \,
\phi_\beta ( {\bf r} ) \nabla_r^2 \phi_\alpha ( {\bf r} ) \; .
\end{equation}
Once $\nabla^2 \phi_\alpha ({\bf r})$ has been evaluated at each grid point
using Eq.~(\ref{eq:finitedifference}), the $T_{\alpha \beta}$
matrix elements are calculated by summing over grid points, just
as for $S_{\alpha \beta}$ (see Eq.~(\ref{eq:overlap})).
In order to evaluate the exchange-correlation, Hartree and pseudopotential
contributions to the total energy, we first need to evaluate the electron
density at each grid point. From Eqs.~(\ref{eq:traces}) and
(\ref{eq:separable}),
the density at grid
point ${\bf r}_\ell$ is:
\begin{eqnarray}
n({\bf r}_\ell) = 2 \sum_{\alpha \beta} \phi_\alpha({\bf r}_\ell)
K_{\alpha \beta}
\phi_\beta({\bf r}_\ell) \; .
\label{eq:density}
\end{eqnarray}
{}From this, it is straightforward to evaluate the exchange-correlation
energy by summing the quantity $n({\bf r}_\ell)
\epsilon_{xc}[n({\bf r}_\ell)]$
over grid points.
The exchange-correlation potential
$\mu_{xc}$ can also be calculated at each point, and is given
as
\begin{eqnarray}
\mu_{xc}({\bf r}_\ell) = \frac{d}{dn} \left\{
n({\bf r}_\ell) \epsilon_{xc}[n({\bf r}_\ell)] \right\} \; .
\end{eqnarray}
To obtain the Hartree energy and potential we use
the fast Fourier transform (FFT) method
to transform the calculated electronic density into reciprocal space, thus
obtaining its Fourier components $\hat{n}_{\bf G}$. The Hartree energy is
then given as
\begin{eqnarray}
E_{\rm H} = 2 \pi \Omega e^2
\sum_{{\bf G} \neq {\bf 0}} | \hat{n}_{\bf G} |^2 / G^2 \; ,
\end{eqnarray}
where $\Omega$ is the volume of the simulation cell.
The Hartree potential in reciprocal space is:
\begin{eqnarray}
\hat{V}_H({\bf G}) = 4 \pi \Omega e^2
\hat{n}_{\bf G} / G^2 \; .
\end{eqnarray}
This can be constructed on the reciprocal-space grid, and transformed
to obtain the Hartree potential in real space. FFT
is, of course, an $O(N \log_2 N )$ operation rather than an $O(N)$
operation, but the difference is negligible for present purposes.
We restrict ourselves here to local pseudopotentials,
so that the value of the total pseudopotential $V_{\rm ps} ( {\bf r}_\ell )$
at grid point ${\bf r}_\ell$ is formally given by:
\begin{equation}
V_{\rm ps} ( {\bf r}_\ell ) = \sum_I v_{\rm ps} ( | {\bf r}_\ell -
{\bf R}_I | ) \; ,
\label{eq:pseudo_sum}
\end{equation}
where $v_{\rm ps} (r)$ is the ionic pseudopotential and ${\bf R}_I$
is the position of ion $I$. In practice, however,
$V_{\rm ps} ( {\bf r}_\ell )$
cannot be calculated like this, because $v_{\rm ps} (r)$ has a
Coulomb tail $-Z |e|^2 /r$ at large $r$, where $Z$ is the core charge.
In order to obtain a linear-scaling algorithm for $E_{\rm ps}$,
we proceed as follows. The ionic pseudopotential is represented
as the sum of the Coulomb potential due to a Gaussian charge
distribution $\eta (r)$ and a short-range potential $v_{\rm ps}^0 (r)$.
The total charge in $\eta (r)$ is $Z |e|$, and the distribution
is given by:
\begin{equation}
\eta (r) = \ Z |e| ( \alpha / \pi )^{3/2} \exp ( - \alpha r^2 ) \; ,
\end{equation}
where the parameter $\alpha$ governs the rate of decay of the
Gaussian. We therefore have:
\begin{equation}
v_{\rm ps} (r) = - \frac{Z |e|^2}{r} \, {\rm erf} ( \alpha^{1/2} r ) +
v_{\rm ps}^0 (r) \; .
\end{equation}
The part of $V_{\rm ps}$ coming from $v_{\rm ps}^0$ can now be
calculated as a direct sum over ions, as in
Eq.~(\ref{eq:pseudo_sum}). Since
$v_{\rm ps}^0$ can be neglected beyond a certain radius, this part
of the calculation scales linearly. The part of $V_{\rm ps}$ coming from
the array of Gaussians can be treated in exactly the same way as
the Hartree potential. The pseudopotential energy is then calculated
by summation over the real-space grid:
\begin{equation}
E_{\rm ps} = \delta \omega \sum_{\ell} V_{\rm ps} ( {\bf r}_{\ell} )
n ( {\bf r}_{\ell} ) \; .
\end{equation}
\subsection{Derivatives and minimization}
Once the contributions to the total energy have been obtained
as outlined above, we need to vary
both $L_{\alpha \beta}$ and $\phi_\alpha$ in order to minimize it.
The $L_{\alpha \beta}$ and $\phi_\alpha$ are independent variables,
and the problem breaks naturally into two separate minimizations that
can be carried out in an alternating manner:
one with respect to $L_{\alpha \beta}$ with fixed $\phi_\alpha$,
and the other with respect to $\phi_\alpha$ with fixed $L_{\alpha \beta}$.
Indeed, the choice of object function can be different
for the two types of variation, and when minimizing with respect to
the $L_{\alpha \beta}$ we find it more convenient to take
$\Omega = E_{\rm tot} - \mu N_e$ as our object function, where $\mu$ is the
chemical potential and $N_e$ is the electron
number. We return to this point below.
Expressions for the derivatives with respect to $L_{\alpha \beta}$
and $\phi_\alpha$ are obtained in Appendix A.
The partial derivative of $\Omega$ with respect to $L_{\alpha \beta}$
is given by
\begin{eqnarray}
\frac{\partial \Omega}{\partial L_{\alpha \beta}} =
\left[ 6(SLH' + H'LS) - 4 (SLSLH' + SLH'LS +H'LSLS)\right]_{\alpha \beta}
\label{eq:partialab}
\end{eqnarray}
where $H' = H - \mu S$, and $H$ is the matrix representation of the
KS Hamiltonian in the support function representation.
It is worth noting that this expression is exactly the same as
would be obtained in a non-orthogonal tight-binding
formalism~\cite{non_orthog_tb}. There is, however, one important
difference: in self-consistent~DFT calculations the Hamiltonian matrix
elements depend on $L_{\alpha \beta}$ through the electronic density
$n({\bf r})$.
The partial derivative of the total energy with respect to $\phi_\alpha$
at grid point ${\bf r}_\ell$ is given by
\begin{eqnarray}
\frac{\partial E_{\rm tot}}{\partial \phi_\alpha ({\bf r}_\ell)} =
4 \delta \omega
\sum_\beta \left[ K_{\alpha \beta} \hat{H} +
3 (LHL)_{\alpha \beta} - 2(LSLHL + LHLSL)_{\alpha \beta}\right]
\phi_\beta({\bf r}_\ell),
\label{eq:partialphi}
\end{eqnarray}
where $\hat{H}$ is the Kohn-Sham operator, which is made to act on
support function $\phi_\beta$.
It is important to notice that because of the spatial localization of
the support functions, and the finite range of $L$,
all the matrices involved in the calculation of these
derivatives are sparse, when the system is large enough.
Provided this sparsity is exploited in the computational
scheme, the method scales linearly with the size of the system.
In the scheme of LNV~\cite{li:nunes:vanderbilt},
it is proposed to work
at constant chemical potential, rather than at constant electron number.
We prefer to maintain the electron number constant.
The variations with respect to $L_{\alpha \beta}$ and
$\phi_\alpha$ will in general cause the electron number to differ from the
correct value, and it is therefore necessary to correct this effect
as the minimization proceeds. We achieve this in the following manner:
during the minimization with respect to $L$, the current search direction
is projected so that it is tangential
to the local surface of constant $N_e$,
i.e. perpendicular to $\nabla_L N_e$ at the current position. This
ensures that the minimization along this direction will cause only a
small change in $N_e$, and it is expected that at the new minimum
$N_e$ will differ only slightly from the required value.
In any case, it is possible to
return to a position as close as desired to the
constant $N_e$ surface by following
the local gradient $\nabla_L N_e$. If the value of the chemical potential
$\mu$
is appropriately chosen, this correction step can be carried out without
losing the reduction in $\Omega$
obtained by performing the line minimization, and
this is why we prefer to take $\Omega$ as the object function instead of the
total energy, when minimizing with respect to $L$. We find that this
scheme is capable of maintaining the electron number close to its correct
value throughout the minimization, and is also simple to implement. The
gradient $\nabla_L N_e$ has elements
\begin{eqnarray}
\frac{\partial N_e}{\partial L_{\alpha \beta}} =
12 (SLS - SLSLS)_{\alpha \beta} \; ,
\end{eqnarray}
which, as all other gradients discussed earlier, can be calculated in
$O(N)$ operations. Minimization with respect to
$\phi_\alpha({\bf r}_\ell)$ will also have
the effect of changing the electron
number. However, given that the two types of variation are performed
alternately, the correction during the $L$ minimization is
sufficient to counteract this effect.
Given that variation of $L_{\alpha \beta}$ causes the electronic
density to change, and this in turn implies that the Hamiltonian
matrix elements change, it would seem necessary to update
the Hamiltonian at each step of the minimization with respect to $L$.
However, we find that this can be avoided by considering $H$ fixed
during this part of the minimization. Strictly speaking, if $H$ is
held fixed while $L$ is varied, we are not minimizing $\Omega =
E_{\rm tot} - \mu N$ but rather $\Omega' = E' - \mu N$, where $E'$ is
given by
\begin{eqnarray}
E' = \mbox{Tr} [( 6 LSL - 4 LSLSL) H] \; .
\end{eqnarray}
If this minimization were carried out through to convergence, this would
be equivalent to diagonalizing $H$ in the representation of the
current support functions. At convergence, it will be found in general
that $L$ and $H$ are not mutually consistent, and if consistency is
required, one needs to update $H$ and repeat the minimization,
iterating this cycle until consistency was achieved. This is not
necessary in practice, because $H$ will be updated at the next
variation with respect to the support functions.
The minimization of $\Omega'$ has practical advantages in that it
avoids the updating of the Hamiltonian at each step, and, because of
its construction, it is a cubic polynomial in every possible search
direction, so it is possible to find the exact location of line minima
during its minimization.
The minimization with respect to $\phi_\alpha({\bf r}_\ell)$ can be carried
out by simply moving along the gradient
$\partial E_{\rm tot} / \partial \phi_\alpha({\bf r}_\ell)$
Eq.~(\ref{eq:partialphi}) (steepest descents)
or by using this expression to construct mutually conjugate directions
(conjugate gradients).
\section{Test calculations}
In order to test our $O(N)$ DFT scheme, we have performed calculations
on a system of 512 Si atoms treated using a local pseudopotential.
The purpose of these tests is to find out how the total energy
depends on the two spatial cutoff radii: the support-region
radius $R_{\rm reg}$, and the $L$-matrix cutoff radius $R_{\rm L}$.
The practical usefulness of the scheme, and the size of system for
which linear-scaling behavior is attained depend on the rate of
convergence of $E_{\rm tot}$ to its exact value as $R_{\rm reg}$
and $R_{\rm L}$ are increased. Here, `exact' refers only to the
absence of errors due to the truncation of $\rho ( {\bf r},
{\bf r}^\prime )$; other sources of inexactness, such as
the use of a discrete grid and a local pseudopotential, are of no
concern here.
The system treated is a periodically repeating cell containing
512 atoms of diamond-structure Si having the experimental
lattice parameter (5.43 \AA). The local pseudopotential
is the one constructed by Appelbaum and Hamann~\cite{appelbaum}, which is
known to give a satisfactory representation of the self-consistent
band structure. The LDA exchange-correlation energy is calculated
using the Ceperley-Alder formula~\cite{ceperley}.
We use a grid spacing of 0.34 \AA,
which is similar to the spacing typically used in pseudopotential
plane-wave calculations on Si, and is sufficient to give
reasonable accuracy. The second derivatives of the $\phi_{\alpha}$
needed in the calculation of $E_{\rm K}$ are computed using the
second-order formula given in Eq.~(\ref{eq:finitedifference}).
A support region is centred on every atom, and each such region contains
four support functions. One can imagine that these support functions
correspond roughly to the single 3$s$ function and the three
3$p$ functions that would be used in a tight-binding
description, but we stress that nothing obliges us to work
with this number of support functions. In keeping with the tight-binding
picture, the initial guess for the support functions is taken to be
a Gaussian multiplied by a constant, $x$, $y$ or $z$, so
that the functions have the symmetry of $s$ and $p$ states. As an initial
guess for the $L$-matrix, we take the quantity $2 I - S$, where
$S$ is the overlap matrix calculated for the initial support
functions. This guess for $L$, which represents the expansion
of $S^{-1} \equiv ( I - ( I - S ))^{-1}$ to first order,
is crude, and does not yield the correct value of ${\rm Tr} \, \rho$.
This error is corrected by displacing $L$ iteratively along the
gradient $\nabla_L N_e$ until $N_e$ is within a required
tolerance of the correct value.
The initial guesses for the $\phi_\alpha$ and the $L_{\alpha \beta}$
define the initial Hamiltonian and overlap matrices. From
this starting point, we make a number of conjugate-gradient
line searches to minimize $\Omega$ by varying $L$, with the Hamiltonian
and overlap matrices held fixed. This is followed by a sequence of
line searches in which the $\phi_\alpha$ are varied. We refer to the
sequence of $L$ moves followed by a sequence of $\phi$ moves
as a {\em cycle}. The entire energy minimization consists
of a set of cycles. In practice, we have found that cycles
consisting of five $L$ moves and two $\phi$ moves
work satisfactorily, and that $E_{\rm tot}$ is converged to within
$10^{-4}$~eV/atom after typically 50-60 cycles. This would not be an
efficient rate of convergence for routine applications, but is more
than adequate for present purposes.
Our test calculations confirm our earlier finding~\cite{hernandez:gillan}
that for the Si perfect
crystal $E_{\rm tot}$ is already quite close to its exact value
when $R_{\rm L}$ = 5.0~\AA. We have therefore used
this value of $R_{\rm L}$ to make calculations of $E_{\rm tot}$ as
a function of $R_{\rm reg}$ (see fig. 1a). The results show that $E_{\rm tot}$
converges very quickly with increasing $R_{\rm reg}$, and that it is
within $\sim$~0.1~eV of its fully converged value for $R_{\rm reg}$ =
3.05~\AA. In order to show how $E_{\rm tot}$ depends on $R_{\rm L}$,
we present a series of results at the two region radii $R_{\rm reg}$ =
2.04 and 2.38~\AA\ (see fig. 1b). These results indicate that there is only a
slow variation with $R_{\rm L}$ and that this variation is almost the same
for different values of $R_{\rm reg}$. This means that it is possible to
converge the total energy to satisfactory accuracy with easily manageable
spatial cutoffs.
It is interesting to know the form of the support functions for the
self-consistent ground state. These are shown in fig.~2 for the case
$R_{\rm reg}$ = 3.05, $R_{\rm L}$ = 5~\AA. The support functions shown
here are the first (initially $s$~Gaussian) and second
($p_x$~Gaussian). Profiles of the support functions along the~[100],
[110] and~[111] directions are shown. The support functions are seen
to be symmetric with respect to the center of the support region ($r =
0$) along the [100] and [110] directions. Along the [111] direction
there is a slight asymmetry resulting from the presence of a nearest
neighbour ion, which lies at 2.35~\AA\ from the origin in the positive
direction. Remarkably, the $s$-like support function seems to be
almost perfectly spherically symmetric, except near the peak at
$r \approx 0.8$~\AA. It is encouraging to see that the support
functions go rather smoothly to zero at the region boundary, and this
confirms that the boundary is having little effect on the results.
\section{Discussion}
We have tried to do three things in this work: to develop the
basic formalism needed to underpin $O(N)$ DFT-pseudopotential
methods; to implement one such method and identify the main technical
issues in doing so; and to present the results of tests on a simple
but important system, which allow us to gauge the usefulness
of the method. We have shown that a rather general class of $O(N)$
DFT-pseudopotential methods can be based on a formulation of DFT in
terms of the density matrix, and that this formulation is equivalent
to commonly used versions of DFT that operate with fractional occupation
numbers. From this viewpoint, the key challenge is to ensure
that the eigenvalues of the variable density matrix lie between 0 and 1,
and we have seen that the method of
LNV~\cite{li:nunes:vanderbilt} gives a way of doing
this. The implementation of the basic ideas has been achieved by
performing all calculations in real space, with the DFT integrals
approximated by sums on a grid -- except for the use of FFT to treat
the Hartree term. An alternative here would be to work with
atomic-like basis functions, but we note that the use of a grid
preserves an important link with conventional plane-wave methods,
as will be analysed in more detail elsewhere. Our test results
on perfect-crystal Si show that the total energy converges rapidly as the
real-space cutoffs are increased, and that it is straightforward
to achieve a precision comparable with that of normal plane-wave calculations.
An important question for any $O(N)$ method is the system size at
which it starts to beat a standard $O(N^3 )$ method -- a plane-wave
method in the present case. This will clearly depend strongly on the system,
but even for Si it is too soon to answer it on the basis of practical
calculations. The cross-over point depends on the prefactor in the
linear scaling, and this is strongly affected by the efficiency of
the coding. All we have attempted to do here is to address the problem
of achieving $O(N)$ behavior. The question of the prefactor is a separate
matter, which will need separate investigation.
It should be clear that there is much more to do before the present
methods can be routinely applied to real problems. We have deliberately
not discussed in detail the problems of doing
calculations on a real-space grid.
Such problems have been discussed outside the linear-scaling context
in several recent papers~\cite{chelikowsky,briggs}, and
it should be possible to apply
the advances reported there to $O(N)$ DFT calculations. In particular,
curvilinear grids~\cite{gygi,zumbach}
for the treatment of strongly attractive pseudopotentials
are likely to be very important for $O(N)$ calculations. We have also not
discussed here the calculation of forces on the atoms, the problems
that may arise when the boundaries of support regions cross grid points,
and the general question of translational invariance
within grid-based techniques.
We have noted already that our method is related to other recently
proposed methods. As shown in sec. 2, the Mauri-Galli-Car scheme is
obtained from ours by taking an alternative polynomial expression
for the density matrix $\rho$ in terms of the auxiliary matrix $\sigma$.
It would clearly be interesting to repeat the calculations done here using
this approach. In a sense, this is what has already been done by Hierse
and Stechel~\cite{hierse:stechel}, except that instead of
performing calculations on a grid,
they use a minimal atomic basis set. The hydrocarbon systems used
by them for test purposes are also rather different from the Si
crystal we have studied.
Finally, we note that our linear-scaling scheme is intended for
calculations on very large systems, and this means that parallel
implementation will play a key role. The test calculations we have
presented were, in fact, performed on a massively parallel machine,
and the parallel-coding techniques we have developed will be described
in a separate paper.
\section*{Acknowledgments}
The work of CMG is supported by the High Performance Computing
Initiative (HPCI) under grant GR/K41649, and the work of EH by
EPSRC grant GR/J01967. The major calculations were done on the
Cray T3D at Edinburgh Parallel Computing
Centre using an allocation of time from the HPCI. Code
development and subsidiary analysis were made using local
hardware funded by EPSRC grant GR/J36266. We gratefully
acknowledge useful discussions with D. Vanderbilt.
|
\section{Introduction}
\setcounter{equation}{0}
The observed correlation function of galaxy clusters and the fluctuation
of microwave background radiation seem to have fractal
structure \cite{LS}, \cite{COB}.
These kinds of observations suggest to attribute a fractal
structure to the universe. When the building blocks of
space-time or some of its subspaces have a fractal structure,
its dimension may have a noninteger value.
The assumption that space has a continuous dimension, was first proposed in
\cite{KMME} where a specific $(d+1)$-dimensional cosmological model with
isotropic and homogenous $d$-dimensional spacelike slices
was proposed.
Its starting point is an extended Einstein-Hilbert Lagrangian in arbitrary
$d$ space dimension
together with a natural constraint between dimension
and scale factor of the universe.
The constraint arises when a cellular structure is attributed to space.
Suppose the dimension of a space $M_0$ is $d_0$ and its size $a_0$.
This space can be constructed from a finite number $N$ of
$d_0$-dimensional cells $e_0$.
Now suppose we have a space $M$ of dimension $d > d_0$. In order to
build such a space from the same number $N$ of cells, these
should have an extra dimension $d-d_0$. Let us take the extra $(d-d_0)$-cell
to be of small size $\ell$ corresponding to some fundamental
length (e.g. the Planck scale).
Then we get the $d$-dimensional volume of $M$ as
$ \mbox{\rm vol} _d(M) = N \mbox{\rm vol} _{d_0}(e_0)\ell^{d-d_0}$,
and analogously the volume of the $d_0$-space $M_0$ as
$ \mbox{\rm vol} _{d_0}(M_0) = N \mbox{\rm vol} _{d_0}(e_0)$.
Sine the volume of a $d$-dimensional space of size $a$ is proportional to
$a^d$, the constraint
$$
({a \over \ell})^d=({a_0 \over \ell})^{d_0}
$$
arises. The model \cite{KMME} predicts that the universe quickly
becomes a FRW universe during its expansion.
Actually this universe may oscillate between a lower scale (the
fundamental length) and an upper scale
(the size of the universe now).
During its expansion, the dimension of
the universe decreases to the present observed value while during
its contraction the dimension of the universe increases to a finite number.
This model solves the horizon problem, since there is no
starting time for the evolution of the universe, and therefore
during several contractions and
expansions all points become correlated. Furthermore
there is no big bang singularity in the model.
On the other hand recent works \cite{BlRZ,Iv} on multidimensional cosmology,
generalizing the Kaluza-Klein idea,
use the possibility to assign further (however constant) dimensions $d_i$
of additional internal factor spaces $M_i$ to the universe
in its early stage of evolution.
Instead of a dynamical reduction of the spacial dimension (like in
\cite{KMME}) here the scales of the internal factor spaces contract.
Multidimensional geometric models
are an interesting class to study in cosmology, because
on one hand, they are rich enough to model features
of phenomenological interest, on the other hand they provide
a well defined minisuperspace. The latter is a convenient
starting point for covariant and conformally equivariant quantization,
with the energy constraint yielding the Wheeler-de Witt (WdW) equation.
Here, we generalize the above works by admitting the factor spaces
to have a fractal dimension which is a smooth function of time.
As an example we study the case
of two factor spaces, one flat and the other compact, where the latter
has a constant dimension.
In fact, the contribution from the scale factor of the compact space
with constant dimension is formally equivalent to some matter field
like the perfect fluid of \cite{KMME}.
Actually, we find it more conclusive (as compared to the standard approach)
to have all matter created from the geometry of space-time.
In Sec. 2 we give the setup of the canonical formalism for
a multidimensional cosmology with Riemannian and, more specifically,
constant curvature factor spaces.
After a proper reformulation
of the Lagrangian and Hamiltonian on the minisuperspace,
the canonical quantization can be applied.
Sec. 3 deals with conformally equivariant quantization on
a minisuperspace.
The first quantization of the energy constraint
is performed in a generally covariant and conformally equivariant
manner. Hence, there is no factor ordering problem.
Sec. 4 then considers the Lagrangian variation with dynamical dimensions,
where a constraint might be taken into account by a Lagrange multiplier.
However the resulting equations of motion
are in general too difficult to be solved analytically.
Therefore in Sec. 5 we consider another, very specific, Lagrangian model
and derive its equation of motion.
In Sec. 6 then the qualitative behaviour of this specific system
is discussed.
Sec. 7 refers to the WdW equation for this system and
Sec. 8 finally resumes the results.
\section{ Riemannian factor space cosmology}
\setcounter{equation}{0}
A convenient reduction of the superspace of geometries is at hand
for the class of multidimensional geometries.
Usually, with $d_i\in \mbox{$\rm I\!N$} $, such a geometry is described by a manifold
\bear{2.8}
M = \mbox{\rm {I\kern-.200em R}}\times M_1 \times\ldots\times M_n, \nonumber \\
D:= \dim M=1+d_1+\ldots+d_n \nonumber \\
g\equiv ds^2 = -e^{2\gamma} dt\otimes dt
+ \sum_{i=1}^{n} a_i^2 \, ds_i^2,
\end{eqnarray}
where $ a_i=e^{\beta^i} $
is the scale factor of the factor space $M_i$ of dimension $d_i\in \mbox{$\rm I\!N$} $,
Here we choose
$$
ds_i^2
=g^{(i)}_{{k}{l}}\,dx^{k}_{(i)} \otimes dx^{l}_{(i)}
$$
such that $ds_i^2$ is a regular bounded measure on $M_i$,
with a finite standard volume
$$
\mbox{\rm vol} _i:=\int_{M_i}ds_i <\infty.
$$
The scale factor exponents ${\beta^i}=\ln a_i$, $i=1,\ldots,n$,
provide a set of coordinates
for the $n$-dimensional minisuperspace $\cal M$ over $M$.
We subject the minisuperspace coordinates $\beta^1,\ldots,\beta^n$
to the principle of general covariance
w.r.t. minisuperspace coordinate transformations.
Like in \cite{BlRZ,Iv}, we restrict here
the $M_i$ to be Einstein spaces of constant curvature.
Then the Ricci scalar curvature of $M$ is
\bear{4.1}
R=e^{-2\gamma}\biggl\{
\biggl[ \sum_{i=1}^{n} (d_i \dot\beta^i) \biggr]^2
+ \sum_{i=1}^{n}
d_i[ { (\dot\beta^i)^2 - 2\dot\gamma\dot\beta^i + 2\ddot \beta^i } ]
\biggr\}
+\sum_{i=1}^{n} R^{(i)} e^{-2\beta^i}.
\end{eqnarray}
The action is usually taken in the standard form
\beq{4.2}
S=S_{EH}+S_{GH}+S_{M},
\end{equation}
where
$$
S_{EH}=\frac{1}{2\kappa}\int_{M}\sqrt{\vert g\vert} R\, dx
$$
is the Einstein-Hilbert action, $S_{GH}$
is the Gibbons-Hawking boundary term,
and $S_M$ some matter term.
Here we choose the boundary conditions such that
the terms with $\dot\gamma$, $\ddot\beta$ from (\ref{4.1}) and $S_{GH}$
cancel out. Since we always have the possibility
to introduce one more dilatonic scale factor from the geometry
instead of some scalar matter field,
here we set $\delta S_M\equiv 0$ without restriction.
Then the variational principle of (\ref{4.2})
is equivalent to a Lagrangian variational principle
over the minisuperspace $\cal M$, given in coordinates $\beta^i$.
$$
S = \int L\, dt,
$$
\bear{4.3}
L = \frac{1}{2}{\mu}
\exp\Bigl\{-\gamma+\sum_{i=1}^{n}d_i\beta^i\Bigr\}
\biggl\{
\sum_{i=1}^{n}{d_i(\dot\beta^i)^2}
- \biggl[\sum_{i=1}^{n}{d_i\dot\beta^i}\biggr]^2 \biggr\} - V(\beta^i)
\end{eqnarray}
with
$$
V(\beta^i) = {\mu}
\exp\Bigl\{\gamma+\sum_{i=1}^{n}d_i\beta^i\Bigr\}
\bigl[ - \frac{1}{2} \sum_{i=1}^{n} R^{(i)} e^{-2\beta^i}\bigr]
$$
where
\beq{4.4}
\mu:=\kappa^{-1}\prod_{i=1}^{n} \mbox{\rm vol} _i.
\end{equation}
Let us define a metric on $ {\cal M} $, given in
coordinates $\beta^i$, $i=1,\ldots,n$.
We set
\beq{4.5}
G_{kl}:=d_k \delta_{kl}-d_k d_l
\end{equation}
$k,l=1,\ldots,n$,
thus defining the tensor components
$G_{ij}$ of the minisuperspace metric
\beq{4.6}
G = G_{ij}d\beta^i\otimes d\beta^j.
\end{equation}
Then with a lapse function $N$, we obtain the Lagrangian
\beq{4.8}
L=\frac {\mu}{2N^2} G_{ij}\dot \beta ^i\dot \beta ^j
-V(\beta^i)
\end{equation}
with the energy constraint
\beq{4.9}
\frac {\mu}{2N^2} G_{ij}\dot \beta ^i\dot \beta ^j
+ V(\beta^i) = 0.
\end{equation}
A convenient gauge for $N$ is the harmonic one \cite{Iv,Rai} given by
\beq{4.7}
N^2:=\exp\biggl\{\gamma-\sum_{i=1}^{n}d_i\beta^i\biggr\} \stackrel{!}{=} 1.
\end{equation}
Nevertheless, here we do not want to restrict to a specific gauge.
Unlike in \cite{BlRZ,Iv,Rai}, we will
prefer to implement a relation like
${\cal C}:=\gamma-\sum_{i=1}^{n}d_i\beta^i=0$ as constraint on the
configuration variables $\beta^i$, rather than as a gauge for $\gamma$.
Note that for a set of $m$ constraints ${\cal C}_k$, $k=1,\ldots,m$
we have to amend the potential
$V\beta^i$ by $-\sum_{k=1}^{m}\lambda_k {\cal C}_k(\beta^i)$, yielding
\beq{4.8c}
L=\frac {\mu}{2N^2} G_{ij}\dot \beta ^i\dot \beta ^j
-V(\beta^i)+\sum_{k=1}^{m}\lambda_k {\cal C}_k(\beta^i),
\end{equation}
with Lagrange multipliers $\lambda_k$.
Let us recall that, although in general the variation of (\ref{4.8c})
w.r.t. $\beta^i$ and $\lambda_k$
does not commute with the implementation of the constraints,
at least the set of solutions for (\ref{4.8}) with
${\cal C}_k(\beta^i)=0$ resolved and substituted before the variation,
is a subset of the solutions of the variation of (\ref{4.8c}) including
the Lagrange multipliers.
Let us consider now the minisuperspace $\cal M$
Its signature of is Lorentzian for $n>1$ (see \cite{Iv}).
After diagonalization
of (\ref{4.5}) by a minisuperspace coordinate transformation
$\beta^i\to \alpha^i$ ($i=1,\ldots,n$), there is just one new
coordinate, say $\alpha^1$, in the direction of which the corresponding
eigenvalue of $G$ is negative. With a further (sign preserving)
coordinate rescaling, $G$ is equivalent to the
Minkowski metric \cite{Rai}. Hence $\cal M$ is flat.
Note that, unlike conformal flatness,
flatness is not an invariant property
under conformal transformation on $\cal M$.
In \cite{Rai} it was pointed out that
while $\beta^i\to\alpha^i$ is only a coordinate transformation
on $ {\cal M} $, it transforms a multidimensional geometry
(\ref{2.8}) with scale exponents $\beta^i$ to another geometry
which is of the same multidimensional type (\ref{2.8}).
This has the same dimensions $d_i$ and first fundamental forms $ds^2_i$,
but new scale exponents $\alpha^i$ of the factor spaces
$M_i$. We can always perform the diagonalization of (\ref{4.5}) such
that $\alpha^1$ and hence $M_1$ belongs to the unique
negative eigenvalue of $G$. This $M_1$ is identified as ''external''
space. The scale factors of the ``internal''
spaces $M_2,\ldots,M_n$ contribute only
positive eigenvalues to the metric of $ {\cal M} $.
$\alpha^1$ assumes in $ {\cal M} $ the role played by time
in usual geometry and quantum mechanics.
In this way the ``external'' space
is distinguished against the ''internal'' ones, since its scale
factor provides a natural ''time'' coordinate on $\cal M$.
Note however that
the ``minisuperspace time'' $\alpha^1$ can be considered
as a time equivalent to $t$ in the underlying multidimensional geometry $g$
only if the space $M_1$ with $\alpha^1$ is
strictly expanding w.r.t. time $t$.
Then the Lorentzian structure of $\cal M$
provides a natural ``arrow of time''
\cite{Ze}.
\section{Canonical minisuperspace quantization}
\setcounter{equation}{0}
Canonical quantization essentially consists in
replacing the constraint equation (\ref{4.9}) by the WdW equation
\cite{ChZ}
\beq{5.1}
\left(-\frac{1}{2}\left[\Delta-\xi_c R\right]+V\right)\Psi = 0
\end{equation}
for a wave function $\Psi$.
We set in the following
\beq{5.2}
N =: e^{-2f}
\end{equation}
and admit $f\in C^\infty(\cal M)$ to be an arbitrary smooth function on
$\cal M$.
In the time gauge given by $f$
the Lagrangian is
\beq{5.3}
L^f:= \frac{\mu}{2} {^{f}}\!G_{ij}(\beta)\dot{\beta}^i\dot{\beta}^j
- V^f(\beta )
\end{equation}
and the energy constraint is
\beq{5.4}
E^f:= \frac{\mu}{2} {^{f}}\!G_{ij}(\beta )\dot{\beta}^i\dot{\beta}^j
+ V^f(\beta ) = 0,
\end{equation}
where
\[
^{f}\!G = e^{2f}G \ \mbox{and} \ V^f = e^{-2f}V.
\]
With the canonical momenta
\beq{5.5}
\pi_{i} = \frac{\partial L^f}{\partial\dot{\beta}^{i}} =
\mu ^fG_{ij}\dot\beta^{j}
\end{equation}
this is equivalent to a Hamiltonian system given by
\beq{5.6}
H^f = \frac{1}{2\mu}(^f\!G)^{ij}\pi_i\pi_j + V^f
\end{equation}
and the energy constraint
\beq{5.7}
H^f = 0.
\end{equation}
The inverse of the minisuperspace metric is given by
$^{f}\!G^{-1} = e^{-2f} G^{-1}$, where for the system with Eq. (\ref{4.5})
the components of $G^{-1}$ are
\beq{5.8}
G^{ij} = \frac{\delta_{ij}}{d_i} + \frac{1}{1-\sum_{i=1}^{n}d_i}.
\end{equation}
At the quantum level $H^f$ has to be replaced by an operator
$\hat{H}^f$, acting in analogy to (\ref{5.7}) as
\beq{5.9}
\hat{H}^f\Psi^f = 0
\end{equation}
on wavefunctions, which are in
a conformal representation of weight $b$ given as
\beq{5.10}
\Psi^f=e^{bf}\Psi.
\end{equation}
Conformally equivariant quantization of $H^f$ from (\ref{5.6}) yields
$$
\hat{H}^f = e^{-2f}e^{bf}\hat{H}\ e^{-bf}
$$
\beq{WdWO}
\hat{H}^f = -\frac{1}{2\mu}\left[ \Delta^f -\xi_c R^f\right]
+{V}^f,
\end{equation}
on wave functions $\Psi^f = e^{bf}\Psi$,
where
\beq{cw}
b=-(n-2)/2,
\end{equation}
\beq{6.14}
\xi_c=\frac{n-2}{4(n-1)},
\end{equation}
\beq{6.3}
\Delta^f=^f\!G^{ij}\nabla^f_i\nabla^f_j
\end{equation}
and both, $R^f$ and the covariant derivative $\nabla^f$, are determined
by the connection $\Gamma^f$ corresponding to the metric ${{}^f}G$.
The WdW equation (\ref{5.9}) is conformally equivariant
if and only if Eq. (\ref{5.9}) for any $f$ is equivalent to
\beq{5.13}
\hat{H}\Psi = 0
\end{equation}
where
\[
\hat{H} = \hat{H}^f\mid_{f=0}\ \mbox{and}\ \Psi =\Psi^f\mid_{f=0}
\]
are the Hamilton operator and the wave function in the gauge $f=0$.
\section {Variation with dynamical dimensions}
\setcounter{equation}{0}
We now pick up the Lagrangian (\ref{4.8c}) of Riemannian factor space
cosmology and consider it as Lagrangian with dynamical dimensions
$d_i$ of some fractal factor spaces.
Hence the dynamical configuration variables
are both $\beta^i$ and $d_i$ for $i=1,\ldots,n$.
In order to implement constraints ${\cal C}_i=0$
on the dynamics of dimensions, we
follow \cite{KMME}. Their constraint is given as follows:
Suppose each factor space is constructed from
a number $P$ of $d$-cells, each of which is a product of
a $d_0$-cell of macroscopic scale and a
$(d-d_0)$-cell of microscopic scale. From these cells we can
built a $d$ dimensional space. For finite $P$ and a vanishing measure on
the $(d-d_0)$-cells,
the $d$-volume of the resulting space is zero. However if the
the $d-d_0$-cells have a length scale $\ell>0$ and nonvanishing
$(d-d_0)$-volume then the volume of each cell is $v_d=v_{d_0} {\ell}^{d-d_0}$.
Now if we write the scale of the factor space as $\ell e^{\beta}$,
we will have
$\ell^d e^{d\beta}= P v_d$ or $e^{d\beta}=e^{d_0\beta_0}$.
Since $d_0$ and $\beta_0$ are constant initial data for the
dynamics (take the present day values), we obtain the constraint
$d\beta=c$ with constant c.
There are many alternatives for generalizing the above considerations
to the case of multidimensional cosmology, e.g.:
1) $\sum_{i=1}^{n}\beta^i d_i= \gamma$
2) Constant $\sum_{i=1}^{m}\beta^i d_i= c_m$ for some $m<n$.
3) Constant $\beta^i d_i= c_i$ for $1\leq i\leq m\leq n$.
Slightly more general than case (1) is the following constraint:
\beq{constr}
\frac{\mu}{2}e^{-\gamma+\sum_{i=1}^{n}d_i\beta^i}={C},
\end{equation}
with a further function $\mu$, and
$C$ independent of the dynamical variables.
This constraint might also be reinterpreted as a generalization
of a harmonic time gauge
of constant $\mu$ and $\gamma$ as in \cite{Rai}.
With the harmonic time gauge many interesting
cosmological models (see also \cite{BlRZ,Iv}) have been constructed.
Note that for constant curvature factor considered here we have
$R^{(i)}=K d_i(d_i-1)$.
Taking the constraint (\ref{constr}),
with Lagrange multiplier $\lambda$,
the Lagrangian (\ref{4.8c}) is
\bear{}
L=\frac{\mu}{2}
\exp\Bigl\{-\gamma+\sum_{i=1}^{n}d_i\beta^i\Bigr\}
\bigl[ G_{ij}\dot \beta ^i\dot \beta ^j
+ K e^{2\gamma}\sum_{i=1}^{n} d_i(d_i-1) e^{-2\beta^i}\bigr]
\nonumber \\
+ \lambda \bigl(e^{-\gamma+\sum_{i=1}^{n}d_i\beta^i}-\frac{2C}{\mu}\bigr).
\end{eqnarray}
We assume that
\beq{volume}
\mbox{\rm vol} _i=m_i(d_i) {\tilde l_i}^{d_i},
\end{equation}
where $\tilde l_i$ is some characteristic length of $M_i$, and
$m_i$ is a function of $d_i$.
For dimension $d$ the coupling $\kappa$ is
\beq{kappa}
\kappa={\ell}^{d-1}
\end{equation}
for some fundamental length $\ell$.
Then with Eq. (\ref{4.4}) and
dimensionless $l_i:=\frac{\tilde l_i}{\ell}$
we obtain
\beq{mu}
{\mu}:=\ell\prod_{i=1}^{n} m_i(d_i) l_i^{d_i}.
\end{equation}
Variation w.r.t. $\lambda$ just reproduces the constraint.
Varying w.r.t. $d_k$ we obtain
$$
0{=}\frac{\partial L}{\partial d_k}
$$
$$
=C
\bigl\{
{\dot \beta^i} {\dot \beta^j} \bigl[
\frac{\partial G_{ij}}{\partial d_k}
+ (\frac{\frac{\partial m_k}{\partial d_k}}{m_k}+\ln l_k+\beta^k ) G_{ij}
\bigr] +
$$
$$
K e^{2\gamma} \bigl[
\sum_{i=1}^{n} d_i(d_i-1) e^{-2\beta^i}
(\frac{\frac{\partial m_k}{\partial d_k}}{m_k}+\ln l_k+\beta^k )+
(2 d_k-1) e^{-2\beta^k}
\bigr]\bigl\} +
$$
\bear{}
\lambda
\{ \beta^ke^{-\gamma+\sum_{i=1}^{n}d_i\beta^i}
+\frac{2C}{\mu}
( \ln{l_k} +\frac{\frac{\partial m_k}{\partial d_k}}{m_k} ) \}.
\end{eqnarray}
The variation w.r.t. $\beta^k$ yields
$$
0{=}{d\over dt}\frac{\partial L}{\partial \dot\beta^k}
-\frac{\partial L}{\partial \beta^k}
$$
$$
=C
\bigl[
\dot G_{kj} \dot \beta ^j + G_{kj} \ddot\beta ^j
\bigr]
+\dot C G_{kj} \dot \beta ^j +
$$
\bear{}
K e^{2\gamma}
2C d_k(d_k-1) e^{-2\beta^k}
- \lambda d^k e^{-\gamma+\sum_{i=1}^{n}d_j\beta^j}.
\end{eqnarray}
For a general space $M_i$ the functions $m_i$ are too
complicated, and the equations above can hardly be solved analytically.
Therefore we have taken the restriction to spaces of constant curvature,
and in the next section we consider a more specific model.
\section {Some specific Lagrangian model}
\setcounter{equation}{0}
Let us now consider more specific cases of the Lagrangian with
$ n $ factor spaces $M_i $ of dimension $ d_i $.
In what follows we specify the Lagrangian for the
cases that all factor spaces are of constant curvature.
Then, a space $M_i$ of positive curvature
is a $ d_i$-dimensional sphere.
For radius $r_i$, its volume is
\beq{spherevol}
\mbox{\rm vol} _i= {{2^{d_i} \pi^{d_i \over 2} }
\over {(d_i -1) \Gamma({d_i\over 2})}}
r^{d_i},
\end{equation}
where $\Gamma $ is the factorial function.
Hence here
\beq{sphere}
m_i(d_i)={ 1 \over {(d_i -1) \Gamma({d_i\over 2})}}
\qquad \mbox{and} \qquad
l_i= 2 \sqrt{\pi} r_i.
\end{equation}
In the case of an open factor
space, we can regularize the measure $ds_i$
such that the volume is
$\int ds_i = \mbox{\rm vol} _i < \infty $.
This could be done e.g. by a conformal
map reducing the radial extension $\infty\to l_i$.
For flat $M_i$ in Eq. (\ref{volume}) $m_i$ is constant.
In the following we assume:
{\hfill \break}
a) One of the factor spaces, say $M_n$ is compact
with constant $R^{(n)}$, all the other factor spaces $M_1,\ldots,M_{n-1}$
are flat.
{\hfill \break}
b) The dimension $d_n$ of this space is
constant, all other dimensions are variable.
{\hfill \break}
c) Here we choose $\gamma=\beta_0 d_1$, where $\beta_0=\beta_1(t_0)$
is the present value of the scale exponent of the external space $M_1$.
With Eqs. (\ref{mu}), (\ref{kappa}) and (\ref{sphere}),
normalizing all volumes with ${\tilde l}_i=\ell \qquad \forall i$,
choosing the constants $m_1,\ldots,m_{n-1}$ such that
$$
\prod_{i=1}^{n-1} m_i = {2\over \ell} (d_n-1) \Gamma({d_n\over 2}),
$$
the Lagrangian (\ref{4.8}) is
\bear{7.3}
L= e^{(\beta_1-\beta_0) d_1+\sum_{i=2}^{n} \beta^id_i}
\big\{ G_{ij} \dot\beta^i \dot\beta^j
+ R^{(n)} e^{2(\beta_0 d_1-\beta^n)}
\big\}
\end{eqnarray}
This is the generalization of \cite{KMME} to multidimensional
vacuum cosmology, here with a potential from the curvature
the compact factor space rather than the potential there.
In the following, we consider the alternative (3) of the previous section,
writing the constraints as
\beq{7.5}
{\beta^i d_i}=c_i
\end{equation}
for $1\leq i\leq m$.
While in the last section, in order to set up the equation of motion,
we had to variate the Lagrangian with respect to $\beta^i$, $d_i$, and
the Lagrange multiplier taking the constraint into account,
here we prefer to resolve the constraints first,
thus reducing the number of variables.
For simplicity we restrict to the case of $n=2$, $m=1$, $k=2$,
assuming that $M_1$ is a flat space of variable dimension $d_1$
subject to a constraint (\ref{7.5}),
and $M_2$ a compact factor space with constant $d_2$.
Then the Lagrangian (\ref{7.3})
simplifies to
$$
L= e^{(\beta_1-\beta_0) d_1+\beta^2 d_2}
\big\{ [-(\dot\beta^1 d_1+ \dot\beta^2 d_2)^2
+(\dot\beta^1)^2 d_1+(\dot\beta^2)^2 d_2]
$$
\bear{Lagr}
+ R^{(2)} e^{2(\beta_0 d_1-\beta^2)}
\big\}
\end{eqnarray}
with
\beq{7.8}
d_1={c\over\beta^1}
\end{equation}
where $c$ is a constant.
Its physical value can be found from the
observational value of the size of the universe \cite{KMME}.
In general, when the constraints are implemented before
the variation of the Lagrangian,
we obtain a subset of the full space of solutions.
Hence, unlike Sec. 4, here we resolve
the different constraints in the very beginning.
Then, besides (\ref{7.8}), the equations of motion will finally be
$$
\ddot\beta^1 +{\beta^1d_2\over {c-\beta^1}}\ddot\beta^2+
{1\over 2}(\dot\beta^1)^2({\beta_0 c\over {(\beta^1)^2}}
-{2c-\beta^1\over {c(c-\beta^1)}})
+ \dot\beta^1 \dot\beta^2 d_2 -
$$
\beq{7.7a}
{1\over 2}\dot\beta^2({d_2(d_2-1)\beta_0 \over {c-\beta^1}}
+{d_2^2\beta^1\over{c-\beta^1}})
+{\beta^1\beta_0\over{2(c-\beta^1)}} R^{(2)}
e^{2(\beta_0{c\over\beta^1}-\beta^2)}
=0
\end{equation}
and
$$
\ddot\beta^2 +{c\over {\beta^1(d_2-1)}} \ddot\beta^1
-{1\over {d_2-1}}(\dot\beta^2)^2+
(\dot\beta^1)^2({c\over{(\beta^1)^2(d_2-1)}}
+({bc\over{(\beta^1)^2}}+d_2){c\over {\beta^1(d_2-1)}})
$$
\beq{7.7b}
+ \dot\beta^2 {1\over {d_2-1}}-\dot\beta^1 {c^2\over{(\beta^1)^2(d_2-1)}}-
({1-{d_2\over 2}}) R^{(2)} e^{2(\beta_0 {c\over \beta^1}-\beta^2)} =0
\end{equation}
\section {Qualitative behaviour of the system}
\setcounter{equation}{0}
{}From the Lagrangian (\ref{Lagr}) the Hamiltonian can be written
\beq{hamil}
H= e^{-\beta_0 d_1+\beta^2 d_2}G_{ij}\dot\beta^i \dot\beta^j -
R^{(2)} e^{\beta_0 d_1+(d_2-2)\beta^2}
\end{equation}
As the Lagrangian of this model is not an explicit function of time, the
Hamiltonian (\ref{hamil}) is a constant of motion.
When the system has a solution $E:=-{H\over {c}}$ is a positive constant.
Let us now first consider the case of constant $\beta^2:=q$.
At the early universe, it is $d_1\gg d_0:=d_1(t_0)$.
we get the following equation
\beq{QU}
(\dot\beta^1)^2+{1\over {c^2}}(\beta^1)^2e^{2\beta_0 d_1-2q} R^{(2)}=
{E\over c^2}({\beta^1})^2 e^{\beta_0 d_1-q d_2}
\end{equation}
Eq. (\ref{QU}) is a $1$-dimensional mechanical system of constant
energy $E$.
The minimum value for $\beta^1$ (where $\dot \beta^1 =0$) is
$$
{\beta^1}_{ \mbox{\rm min} }={c\beta_0 \over {\ln({E\over R^{(2)}}})-q(d_2-2) }.
$$
Therefore a maximum value for the dimension of the early universe can
be derived
$$
{d_1}_{ \mbox{\rm max} }={{ {\ln({E\over R^{(2)}}})-q(d_2-2) } \over \beta_0}
$$
On the other hand from \ref{QU}, one finds that
$$
\ddot \beta^1 > 0 \qquad \forall t
$$
Therefore $\beta^1_{min}$ is a local minimum. As a result the universe of
this model contracts to a minimum and expands again. This behaviour
is similar to the previous work \cite{KMME}.
The behaviour of this universe at larger $\beta^1$
will be determined by the factor
$(\beta^1)^2e^{\beta_0 c\over {\beta^1}}$ in equation (6.2). Therefore
with $k=\pm \sqrt{{E\over R^{(2)}}e^{-q(d_2-2)}-1}$,
$\beta^1$ will behave as $e^{kt}$.
Therefore this model predicts the late expansion of the external space
to be inflationary.
In the case $\dot \beta^2\neq 0$,
for the early universe $\beta^1_{ \mbox{\rm min} }$
is shifted to a larger value. Thus
the dynamics of the internal factor space $M_2$ prevents the external
factor space $M_1$ (our universe) from reaching a
singularity in the very early universe.
Another virtue of the second factor space to control the inflation.
The graceful exit of the
factor space $M_1$ can only be obtained as an effect of $\beta^2(t)$.
Complementary to the first case above,
we have to examine also the situation for constant $\beta^1$.
There, the behaviour of $\beta^2$ can be solved
easily from (\ref{QU}). For $d_2=2$ we find $\beta^2$
proportional to $\ln(t)$.
Therefore the scale factor of the second space decreases as ${1\over t}$.
As a result the general behaviour of the continuous dimensional space
is an exponential expansion and the second compact constant dimensional
space contracts slowly. An exact analysis of the behaviour , needs
the simulation of (\ref{7.7a}) and (\ref{7.7b}).
\section{WdW equation for the model}
\setcounter{equation}{0}
Recall that the
Lagrangian in an arbitrary gauge, with $N=e^{-2f}$, is given as
$$
L^f:=NL= {1\over N}G_{ij}\dot\beta^i\dot\beta^j-NV(\beta^i)
$$
\beq{Lf}
={1\over N}[G_{ij}\dot\beta^i\dot\beta^j-N^2V(\beta^i)].
\end{equation}
With the Lagrangian (\ref{Lagr}) and the
gauge $f={1\over 4}(\beta_0 d_1 -\sum_i \beta^i d_i)$
we get
\beq{Lag}
L^f=e^{{1\over 2}((\beta^1-\beta_0) d_1 + \beta^2 d_2)}
[G_{ij}\dot\beta^i\dot\beta^j
+ R^{(2)} e^{2(\beta_0 d_1-\beta^2)}].
\end{equation}
Now we consider the corresponding Hamiltonian
$$
H^f:=NH={1\over N}[G_{ij}\dot\beta^i\dot\beta^j+N^2V(\beta^i)],
$$
and change from the gauge
$f=-{1\over 4}(\beta_0 d_1 -\sum_i \beta^i d_i)$ to $f=0$.
Then, in the new gauge the Hamiltonian is
$$
H^0=G^{ij}\dot\beta^i\dot\beta^j+V(\beta^i)
$$
\beq{H0}
={1\over 4}G^{ij}\pi_i \pi_j+ V(\beta^i)
\end{equation}
where $V(\beta^i)=-R^{(2)}e^{2(\beta_0 d_1-\beta^2)}N^{-2}=
-R^{(2)}e^{\sum_i \beta^i d_i +\beta_0 d_1-2\beta^2}$.
Canonical quantization in this gauge yields
\beq{WdWop}
\hat H^0=-{1\over 4}\Delta +V,
\end{equation}
$$
\Delta=G^{ij}\nabla_i\nabla_j={1\over \sqrt{-\det G}}
\frac{\partial}{\partial \beta^i}
(\sqrt{-\det G}G^{ij}
\frac{\partial}{\partial \beta^j}).
$$
Recall that here $n=2$, hence Eq. (\ref{WdWO}) applies simply with
$\xi_c=0$ and $b=0$.
With $d_1={c\over \beta^1}$ and constant
$d_2$ the inverse of $G_{ij}$ is
\beq{invG}
(G^{ij})=
\left [\begin {array}{cc}
{\beta^1\over c}+{c \over c(1-d_2)-\beta^1} &
{c \over c(1-d_2)-\beta^1} \\\noalign{\medskip}
{c \over c(1-d_2)-\beta^1} &
{1\over d_2}+{c \over c(1-d_2)-\beta^1}
\end {array}\right
],
\end{equation}
and $\sqrt{-\det G}={1\over{\beta^1}}\sqrt{c(d_2-1)d_2\beta^1+c^2d_2} $.
Since with $n=2$ the conformal weight (\ref{cw}) is $b=0$,
a solution $\Psi=\Psi^0$ of the WdW equation $\hat{H}^0\Psi=0$
in the gauge $f=0$ is also a solution of the
the WdW equation $\hat{H}^f \Psi^f = e^{-2f}\hat{H}\Psi=0$
in the original gauge $f={1\over 4}((\beta_0 - \beta^1) d_1-\beta^2 d_2)$.
Actually here we have minisuperspace curvature $R[G]=0$ and
$\det G<0$ for $\beta_1>0$ and $d_2>1$.
Hence the minisuperspace $\cal M$ is
the $2$-dimensional flat Minkowski space,
like in the analogous case of constant dimensions.
In order to get a feeling for the qualitative structure of the WdW equation,
let us write it explicitly along a line of constant $\beta^2=q$.
There, at the limit of large $d_1$ it is
\beq{WdW2}
\biggl\{ (\frac{\partial}{\partial \beta^1})^2
+ {c\over {(\beta^1)^2}} \frac{\partial}{\partial \beta^1}
+ 4K d_2 (d_2-1)^2e^{c+q(d_2-2)}e^{\beta_0 c\over \beta^1} \biggr\}
\Psi\vert_{\beta^2=q}(\beta^1)=0
\end{equation}
Orthogonally to this, the WdW equation
along some line of constant $\beta^1=\frac{c}{d_1}=p$ is
\beq{WdW1}
\biggl\{
\frac{d_1-1}{d_2 (d_1 d_2- d_1 +1)} (\frac{\partial}{\partial \beta^2})^2
- 4 Kd_2(d_2-1)e^{c+\beta_0 d_1}e^{\beta^2(d_2-2)}
\biggr\} \Psi\vert_{\beta^1=p}(\beta^2)=0.
\end{equation}
Finally we want to compare our previous minisuperspace to the case
with $2$ constraints
$d_i\beta^i=c_i$, for {\em both}, $i=1$ {\em and} $i=2$.
There
\beq{invG2}
(G^{ij})= (\beta_1\beta_2-\beta_1 c_2-\beta_2 c_1)^{-1}
\left [\begin {array}{cc}
(\beta^2-c_2){\beta^1\over c_1} &
{\beta^1\beta^2} \\\noalign{\medskip}
{\beta^1\beta^2} &
(\beta^1-c_1){\beta^2\over c_2}
\end {array}\right
],
\end{equation}
and $\sqrt{-\det G}={1 \over\beta^1\beta^2}
\sqrt{c_1 c_2 (c_2\beta_1+c^1\beta^2-\beta^1\beta^2)} $.
Here, for $c_{1/2}>0$ and $d_{1/2}>1$, the minisuperspace
$\cal M$ is always Lorentzian. However, unlike the previous
example, it is no longer homogeneous, since
\beq{R2}
R[G]=-\frac{1}{2}\beta_1\beta_2/(c_2\beta_1+c^1\beta^2-\beta^1\beta^2)^2.
\end{equation}
For $\beta_1\to\infty$ or $\beta_2\to\infty$ the curvature decays
$R[G]\to 0$ and $\cal M$ becomes the usual homogeneous
Minkowskian space.
For $\beta_1\to\ 0$ or $\beta_2\to 0$ there appears a singularity
in the minisuperspace curvature, $R[G]\to \infty$ .
However, taking the conformal quantization scheme
seriously, we should be aware that the minisuperspace curvature
itself is not an invariant property of the quantum system
because it may be changed by conformal transformations.
More specifically, in our case the minisuperspace $\cal M$ is
$2$-dimensional, hence there exists a gauge $f$ such that
$\cal M$ is flat and therefore also homogeneous.
So for this model the inhomogeneity (\ref{R2})
is only a property for the specific gauge $f=0$.
According to \cite{Rai}, in canonical minisuperspace quantization
$\beta^1$ and $\beta^2$ are just coordinates for $\cal M$.
Hence the singularity of (\ref{R2}) is the analogue
of a classical coordinate singularity.
\section{Conclusion}
\setcounter{equation}{0}
We have investigated the effect of dynamical dimensions
of fractal factor spaces
on the evolution of a multidimensional cosmological model.
In a mathematically closed approach, we could have
set up the differential geometry of fractal spaces in the
very beginning. This can essentially be done using a
definition of generalized manifolds by simplicial complexes
(as exemplified e.g. in \cite{Hoef}).
However, for brevity,
here we rather preferred to derive the Lagrangian as for constant dimensions,
and then to consider the dimensions as variables of just this Lagrangian.
More specifically, we discussed a multidimensional cosmological model with
two factor spaces, one of them flat with dynamical dimension, the
other compact with constant curvature and constant dimension.
In fact the latter behaves like a matter field.
By qualitative analysis, the behaviour of the system
shows that the universe, i.e. the factor space $M_1$,
contracts, passes through a state of minimum size
(maximum dimension), and expands.
For a static internal space (compare \cite{BlRZ}),
i.e. for constant $\beta^2=q$,
in the late expansion, for large $\beta^1$, the universe inflates double
exponentially. Therefore, the dimension decreases very fast to its
minimum value.
Actually, the scale factor $\beta^2$ of the
compact constant dimensional space $M_2$ controls the
behaviour of the expansion of universe.
Eventually the dynamics of $\beta^2$ might be the only way
to obtain a graceful exit to the effective model \cite{KMME}.
Again this model, as in the case of \cite{KMME}, has no big bang
singularity.
However, in order to find the complete behaviour of this
model, a more sophisticated analysis would be required.
In the case of a static internal space $M_2$,
near the maximun of $d_1$, i.e. minimum of $\beta^1$, the present model is
effectively represented by a model \cite{BlRZ,Rai} with only constant
dimensions. There the minimum value of $\beta^1$ could be related
to the quantum creation of the real Lorentzian space-time
{}from an Euclidian region. So, the present model is compatible with a
quantum creation of our universe.
We have further derived the WdW equation for the
minisuperspace of this model. Like in previous works on multidimensional
cosmology, here the metric describes a Minkowskian minisuperspace.
For the slightly more general case of both factor spaces
subject to the same type of constraint (\ref{7.5}), in the gauge
$f=0$, we find an inhomogeneity and singularities in
the minisuperspace curvature $R$. However, in the case of only $2$
factor spaces, the conformal class of $G$ is in any case the flat
Minkowskian one.
{\hfill \break} \nl
{\bf Acknowledgement}
{\hfill \break}
{
Support by DFG grant Bl 365/1-1 and Schm 911/6-2
is gratefully acknowledged.
M. M. thanks for hospitality at the Projektgruppe
Kosmologie of Universit\"at Potsdam.
}
|
\section{Table Caption}
\begin{center}
\begin{tabular}{|| c|c|c|c|c|c|c|c|c ||}\hline
\squeezetable
$\lambda$
& 0.0005 & 0.001 & 0.005 & 0.01 & 0.05 & 0.1 \\
\hline
$\kappa^{LW} _{crit}$
& 0.125101 & 0.125202 & 0.125991 & 0.126968 & 0.132368 & 0.13601 \\
\hline
$\alpha$
& 0.99997 & 0.99993 & 0.99972 & 0.9993 & 1.0073 & 1.0275 \\
\hline
\end{tabular}
\end{center}
The critical points $\kappa_{crit}$ versus $\lambda$.
$\kappa_{crit} ^{LW}$ is taken from Ref.\cite{Luscher}.
$\alpha := \kappa^{KS} _{crit}/ \kappa^{LW} _{crit}$ denotes the
ratio between the results of this work and Ref.\cite{Luscher}.
In this work, $\kappa_{crit}$ has been determined by the condition
that the renormalised mass $m_R$ becomes imaginary. $\Lambda/\Delta p = 4$.
\end{document}
|
\section{Introduction}
The question of non-local quantum correlations versus local realism,
first raised in the famous EPR paper \cite{epr}, has held the interest
of the physics community since. J. S. Bell \cite{bell} showed that
the predictions of quantum mechanics are incompatible with any model
based on local realism. The pioneering experimental work of A. Aspect
{\it et al.} \cite{aspect} and others \cite{mandel}
supports the
predictions of quantum mechanics and contradicts local realism:
Bell inequalities applicable to the various experimental arrangements
have been shown to be violated. It should be mentioned that
some aspects of the
experimental setups have been criticized and questioned
\cite{santos}. Problems of experimental
bias or enhancement of particular polarization states by detection
systems were experimentally checked by T. Haji-Hassan {\it et al.}
\cite{hassan}
and found absent. And more recently Kwiat {\it et al.} \cite{kwiat} have
proposed and described an experimental arrangement that overcomes
shortcomings of previous experiments. While experiments are still open
to criticism, it is generally accepted that local realism
is untenable. In this Letter we
assume that in nature there exist non-local
correlations, as predicted by quantum mechanics,
and we address the following
question: Can an experimenter {\it non-locally tamper}
with non-local correlations, without violating relativistic causality?
Quantum mechanics predicts non-local correlations; however, it does
not provide an ``explanation" about what creates them.
Several theoretical
models go beyond quantum mechanics and propose to explain
the phenomenon of non-local correlations via a superluminal
``communication link'' \cite{bohm}. If one accepts the possibility of
a communication link, then a natural next step would be to
probe whether it is possible to tamper with this link and {\it jam}
the superluminal communication \cite{shimony}.
Up to now, the possibility of jamming non-local
correlations has not received due consideration, perhaps because of
a tacit assumption that such tampering necessarily violates
relativistic causality.
(The expression {\it relativistic causality} is used here to
denote the principle that information cannot be transferred at speeds
exceeding the speed of light.)
In this Letter we show that jamming of
non-local correlations can be consistent with relativistic
causality. Our results are independent of the model used to describe
how the non-local quantum correlations arise, that is, the
nature of the
superluminal communication link, and they apply to any jamming
mechanism.
\section{The Jamming Scheme}
Jamming might take many forms. The following discussion does
not discuss a mechanism for jamming; rather, it defines the
constraints that any jamming mechanism
must obey in order to be consistent with relativistic causality.
In order to derive and
illustrate the constraints, it is convenient to consider a particular
experimental arrangement which can be subjected to jamming
\cite{generic}.
We will
consider an EPR-Bohm experimental arrangement
to study pairs of spin-$1/2$
particles entangled in a singlet state \cite{eprb}. Spacelike separated
spin measurements on these pairs allow a test of the Bell inequalities.
Suppose that two experimenters, Alice and Bob, perform
the spin measurements. One particle of each entangled pair arrives at
Alice's analyzing station and the other particle arrives at Bob's.
When Alice and Bob get together and combine the results of their
measurements, they will find violations of the Bell inequalities,
as predicted by quantum mechanics \cite{bell}.
We now introduce a third experimenter, Jim, the jammer, who has
access to a jamming device which he can activate, at will, and tamper
with
the communication link between each entangled pair of particles.
His action is spacelike separated from the measurements of
Alice or Bob or from both of them.
Jamming acts at a distance to modify the correlations between
the particles; it disturbs
the conditions which make possible the phenomenon of
non-local quantum correlations. Therefore, the correlations
measured jointly by Alice and Bob will not agree with the predictions
of quantum mechanics.
Jamming is truly non-local and cannot be carried out within the
framework of
quantum mechanics. For example,
consider three systems, $S_1$, $S_2$ and $S_3$,
in a quantum state $\Psi_{123}$. Let experimenters near
$S_1$ and $S_2$ measure $A^{(1)}$ and $A^{(2)}$, with eigenstates
denoted by
$\vert a^{(1)}_i \rangle$ and $\vert a^{(2)}_j \rangle$, respectively.
The only freedom available to an experimenter near
$S_3$ is the choice of what local operator
$A^{(3)}$ to measure. But the probabilities $P(a^{(1)}_i ,a^{(2)}_j )$
for
outcomes $A^{(1)} =a^{(1)}_i$ and $A^{(2)} =a^{(2)}_j$,
\begin{equation}
P(a^{(1)}_i ,a^{(2)}_j )
=\sum_k \vert \langle \Psi_{123} \vert a^{(1)}_i
,a^{(2)}_j ,a^{(3)}_k \rangle
\vert^2~,
\end{equation}
are {\it independent} of the choice of operator $A^{(3)}$. Thus
no measurement on $S_3$ can affect the results of the measurements
performed on $S_1$
and $S_2$, even if the three systems have interacted in the past
\cite{note}.
In general, jamming would allow Jim to send superluminal signals.
The constraints that
must be satisfied in order to insure that
Jim cannot send superluminal signals are embodied in two conditions.
The first condition,
the {\it unary condition}, a necessary but
not sufficient condition, requires that Jim not be able to send
signals
to Alice or Bob {\it separately}. In effect
this condition demands that
Alice and Bob, separately, measure
zero average spin along any axis. Explicitly, let $N_a (+)$ and
$N_a (-)$ tally the number of spin-up and spin-down results,
respectively,
found by Alice for a given axis. For the same axis, let $n(k,l)$ tally,
in the absence of jamming,
the joint results of Alice and Bob. The parameters
$k$ and $l$ denote, respectively, the results
( $+$ or $-$ ) of the
polarization measurements
carried out by Alice and Bob.
Let $n^\prime (k,l)$ tally, in the presence of jamming, the
corresponding polarization measurements carried out by Alice and Bob.
The unary condition imposes the
following relations between $n(k,l)$ and $n^\prime (k,l)$:
\begin{eqnarray}
\label{Na}
N_a (+) &=& n(+,+) + n(+,-) = n^\prime (+,+) + n^\prime (+,-)\nonumber \\
N_a (-) &=& n(-,+) + n(-,-) = n^\prime (-,+) + n^\prime (-,-).
\end{eqnarray}
A similar set of relations holds for the results $N_b (+)$ and
$N_b (-)$ found by Bob. Hence regardless of
whether Jim has activated the
jamming device, Alice and Bob will find that the average spin projection
along any axis tends to zero, and Jim cannot send superluminal signals,
separately, to either Alice or Bob.
The unary condition
allows a range of possibilities for the
jammed correlations: from
correlations which are only slightly different from those predicted
by quantum mechanics, down to completely random
correlations. In particular, the unary condition allows
conservation of angular momentum, {\it i.e.} perfect anticorrelation
of spin components along any parallel axes.
\section{The Space-Time Window}
As stated in the previous section,
the unary condition is a necessary but not sufficient condition. For
jamming to respect relativistic causality, we must also restrict the
relationships in space and time among the three events $a$, $b$ and $j$
generated, respectively, by Alice, Bob and Jim. Fig. 1
shows the geometry of
three different configurations of an EPR-Bohm experimental setup along
with the corresponding Minkowski diagrams of the events
$a$, $b$ and $j$. In the configuration shown in
Fig. 1(a), jamming is
{\it not} permitted. Here Alice and Bob are in close proximity while
Jim is far away. If jamming were permitted, Alice and Bob could
---immediately after Jim activates the jamming device---measure the spin
projections of their respective particles and combine their results
to determine the spin correlations. They would find spin correlations
differing from the predictions of quantum mechanics and infer that
Jim activated the jamming device. The corresponding Minkowski
diagram, Fig. 1(b), shows that
the future light cones of $a$ and $b$ overlap,
in part, outside the future light cone of $j$. A light signal
originating at $j$ cannot reach this overlap region
of $a$ and $b$,
where Alice and Bob can combine their results. Were jamming
possible here, it would violate relativistic causality.
Fig. 1(c) shows a configuration that would also permit
superluminal signalling: Jim obtains the results of Alice's
measurements prior to
deciding whether to activate the jamming device. Bob is far from both
Alice and Jim. The corresponding Minkowski diagram, Fig. 1(d), shows
that $a$ precedes $j$ by a timelike interval and both $a$ and $j$ are
spacelike separated from $b$. Since Jim has access to Alice's results,
he can send a superluminal signal to Bob by {\it selectively} jamming:
For instance, suppose Jim activates the jamming device only when Alice
obtains the value $ + 1/2$ for the projection of the
spin of a particle. Bob will, then, find that
the average spin component along a given axis does {\it not} tend to
zero. The preceding can be demonstrated
by comparing the results of the spin measurements
$N_b (+)$ and $ N_b (-)$, carried out by Bob in the absence of jamming,
Eqs. (3), and in the presence of selective jamming, Eqs. (4).
The notation
previously defined is used in Eqs. (3-4).
\begin{eqnarray}
\label{Nb}
N_b (+) &=& n(+,+) + n(-,+) \nonumber \\
N_b (-) &=& n(+,-) + n(-,-) ~,
\end{eqnarray}
\begin{eqnarray}
\label{sel}
N_b (+) &=& n^\prime (+,+) + n(-,+) \nonumber \\
N_b (-) &=& n^\prime (+,-) + n(-,-)~.
\end{eqnarray}
Hence the results obtained by Bob in the presence of selective
jamming will be different from those obtained in the
absence of jamming unless
$n^\prime(+,+) =n(+,+)$ and $n^\prime(+,-) =n(+,-)$. However, the
latter requirements imply
that jamming, in this configuration, can not have any
discernible effect, i.e. jamming in this configuration is
impossible.
To eliminate configurations which
allow violations of
relativistic causality, as shown in Fig. 1(a) to Fig. 1(d),
we further restrict jamming by
imposing a second condition, the {\it binary
condition}. The binary condition, which is manifestly covariant,
demands that the overlap of the future light cones
of $a$ and $b$ lie entirely within the future light cone
of $j$ and therefore a light signal emanating from $j$ can
reach the overlap region.
The configuration shown in Fig. 1(a) and 1(b), which allows an overlap of
the future light cones of $a$ and $b$ outside of the future light
cone of $j$, is therefore forbidden. The configuration shown in
Fig 1(c) and 1(d), a configuration for
selective jamming, violates the unary condition
and it is also disallowed by the
\bigskip
\begin{figure}
\centerline{\epsfig{file=fig89ok.eps}}
\label{fig:hyp}
\caption{The geometrical configurations showing the
source $S$ of pairs of quantum systems, the jammer $J$, and the
experimenters Alice $A$, and Bob $B$.
(a) A and B are close to each other while J is far
from both of them.
(c) A and J are close to each other while B is far
from both of them.
(e) A, B and J are all far from each other; J is stationed near the
source and A and B are at opposite ends of an EPR-Bohm setup.
Corresponding Minkowski diagrams showing the events
$a$, $b$ and $j$. (b) The future light cones of $a$ and $b$ have some
overlap outside the future light cone of $j$.
(d) A possible configuration for selective jamming.
(f) A configuration satisfying the binary condition. The
future light cones of $a$ and $b$ overlap only within
the future light cone of $j$.}
\end{figure}
binary condition. A configuration which satisfies the binary condition
is shown in Fig. 1(e) and 1(f).
The constraints to which a jamming configuration must conform,
in order not to violate relativistic causality,
are embodied in the unary and binary conditions.
These conditions are manifestly Lorentz
invariant. However, the time sequence of the events $a$, $b$ and $j$ is
not. A time sequence $a$, $j$ and $b$ in one Lorentz frame may
transform into $b$, $j$ and $a$ in another Lorentz frame.
Hence while one observer will claim that Alice completed
her measurements before Jim activated his
jamming mechanism and thus
Jim affected only the results of Bob's measurements, another observer
will claim that Bob carried out his measurements
first and Jim affected only Alice's results. Similar situations are
encountered in quantum mechanics where different observers in
different Lorentz frames will give conflicting interpretations
of the same set of events. For
example, with respect to an entangled pair of particles in an EPR-Bohm
experiment, the question of which observer caused the collapse of the
entangled state has no Lorentz-invariant answer \cite{aa}.
If jamming is possible then one must accept the possibility of
reversal of the {\it cause-effect} sequence \cite{dbohm};
however, the allowed configuration
which satisfies the {\it unary} and {\it binary} conditions does not
lead to contradictory causal loops, i.e. no {\it effect} can send a
signal to its {\it cause}.
Indeed, consider one
jammer, $J$, who acts on the correlations between two spacelike separated
events, $a$ and $b$.
We first recall that the
unary condition precludes signalling to $a$ and $b$, separately,
by $j$; therefore, only the combined results of the measurements
of $a$ and $b$ can reveal whether $J$ activated a jamming mechanism.
In order to complete a contradictory casual
loop one must gather the
results of the measurements of $a$ and $b$ into the past light cone of
$j$ and then send a signal to $j$, the {\it cause}.
But the binary condition requires that the overlap of the future
light cones of
$a$ and $b$ be completely contained in the future light cone of
$j$, so the
only place where information from $a$ and $b$ can be
put together by means of
ordinary signals is the future of $j$. One might suppose that
other jammers, using their non-local action, could somehow transmit the
information from $a$ and $b$ into the past light-cone of $j$.
Such a scheme would
require at least two
more jammers. Since these jammers must have access to the results of
$a$ and $b$, we place $j_1$ and $j_2$ (generated by
$J_1$ and $J_2$) at timelight separations,
respectively, from $a$ and $b$. Events $a$ and $b$ are spacelike
separated from each other and from $j$, so $j_1$ and $j_2$
will either be spacelike separated from $j$ or in its future
light cone.
The cases of $J_1$ and $J_2$ are similar, so we discuss only
$J_1$; however, the conclusions reached apply equally to
$J_1$ and $J_2$. The jammer,
$J_1$, can communicate the results of $a$ by jamming or not jamming
the non-local correlations between pairs of entangled particles
measured at events $a_1$ and $b_1$. Notice that in
order to communicate the result of a single measurement
done at $a$, $J_1$ must jam (or not jam) an
ensemble of EPR pairs. The result of a single measurement
carried out at $a$ is recovered from the correlations
determined from the combined measurements done at $a_1$ and $b_1$.
For the jammer
$J_1$ to gather the information at $a$ into the past light cone of
$j$ requires that both
$a_1$ and $b_1$ lie in the past light cone of $j$, i.e. $j$ lies in
the overlap of the future light cones of $a_1$ and $b_1$.
This requirement, however, is incompatible with the
binary condition when applied to the triplet of events,
$a_1$, $b_1$ and $j_1$, which requires that the overlap of
$a_1$ and $b_1$ be contained within the future light cone of $j_1$.
This, in turn, implies that
$j$ will lie in the future light cone of $j_1$,
contradicting the assumption that $j_1$ is either
spacelike separated from $j$ or in $j$'s future light cone.
Consequently, at least one event
$a_1$ or $b_1$ must be spacelike
separated from $j$. Therefore the introduction of $J_1$ does not
help to gather the results of $a$ into the past light cone of $j$.
Then, by induction, we find that no scheme
to close a contradictory causal loop, by introducing
any number of jammers, can succeed.
\section{Conclusions}
In quantum mechanics
non-local correlations are well established; however, these correlations
cannot be used to send superluminal signals. In this Letter we have
raised the question of whether a form of non-locality beyond quantum
mechanics---non-local tampering with quantum correlations---could also
respect relativistic causality.
We find that jamming configurations which obey
two conditions---the {\it unary} condition, which forbids
superluminal signalling to either of two
experimenters, and the {\it binary} condition, which restricts the
space-time configuration of the two experimenters and the
jammer---respect relativistic causality. For these configurations,
the cause-effect sequence might not be preserved in all
Lorentz frames; however, they do not
lead to contradictory causal loops.
Hence, we find that a stronger form of non-locality
than that arising in quantum mechanics---action at a distance
rather than non-local correlations---is consistent with relativistic
causality. \cite{shimony,axiom,futur}
The results presented in this Letter are independent of the model used
to describe the nature of the non-local correlations and apply
to any jamming mechanism. Experimental studies, to date,
have not tested the possibility of jamming.
We suggest that current and projected EPR-Bohm experiments
test the possibility of jamming in configurations consistent with the
constraints derived in this Letter.
The constraints on the jamming configuration, however, because
of their generality, do not themselves suggest a preferred mechanism
for carrying out the jamming procedure.
We thank Y. Aharonov for helpful discussions. The research of D. R.
was supported by the State of Israel, Ministry of Immigrant Absorption,
Center for Absorption in Science.
|
\section{Introduction}
The minimal supersymmetric standard model (MSSM) is the
most attractive candidates for the realistic theory
beyond standard model.
The naturalness problem is elegantly solved by the introduction
of supersymmetry (SUSY)\cite{Naturalness}.
The SUSY requires new particles called $\lq$superpartners'.
Those masses are free parameters in the MSSM,\footnote{
In this paper, we do not assume the universality on the
soft SUSY breaking parameters from the beginning
when we use the terminology $\lq$MSSM'.} but are estimated
as at most order 1 TeV from the naturalness argument.
The search for $\lq$superpartners' is one of the main purpose
in the experimental projects by the use of huge colliders,
which have been planed now\cite{JLC}.
It is, however, believed that the MSSM is not the ultimate theory
because there are many problems not to be solved in it.
Here we pick up two problems.
First there are so many free parameters
to be fixed only by experiments for the present.
In addition to gauge couplings and Yukawa couplings,
soft SUSY breaking parameters appear, i.e., gaugino masses,
scalar masses and scalar trilinear couplings are
all arbitrary ones.
Hence the MSSM lacks predictability.
Second the mechanism of SUSY breaking is unexplained.
This problem is partly related to the first one since
the pattern of soft SUSY breaking terms
depends on the SUSY breaking mechanism.
It is expected that they are solved in more fundamental
theory.
Supergravity theory (SUGRA)\cite{SUGRA} is
an attractive candidate.
When we take SUGRA as an effective theory
at the Planck scale $M_{Pl}$,
SUGRA has an interesting solution.
There exists such a scenario\cite{Hidden}
that the SUSY is spontaneously broken
in the so-called hidden sector
and the effect is transported to the observable sector
through the gravitational interaction.
As a result, the soft SUSY breaking terms appear
in our visible sector.
The form of soft SUSY breaking terms
is determined by the structure of SUGRA.
A simple choice is a theory such that
soft parameters take universal values at the
gravitational scale $M \equiv M_{Pl}/{\sqrt {8\pi}}$, e.g.,
the scalar potential derived from the SUGRA
with a minimal K\"ahler function has
the universal scalar mass $m_0$ and
the universal scalar trilinear coupling constant $A$.
Those values at low energy are calculated by using
renormalization group equations.
The analyses based on the MSSM are energetically
investigated\cite{MSSM}.
Most of them are highly constrained by the assumption
that the soft SUSY breaking parameters are universal at $M$
or a unification scale $M_X$.
This assumption is quite interesting
because the theory has high predictabilities
and is testable enough, but it is difficult to say that this
type of approach is completely realistic.
Let us describe the reason why the universality at $M$ is
not necessarily realistic.
First the assumption of the universal scalar mass is motivated
by the fact that the flavor-changing neutral current
(FCNC) processes are suppressed experimentally\cite{FCNC}.
However, we can relax this assumption since the suppression
of FCNC processes due to SUSY particle loops requires only
the degeneracy among squarks with a same flavor.
Second there is no strong reason that the realistic SUGRA takes
the minimal structure.
In fact, the effective SUGRAs derived from superstring theories
(SSTs) have, in general, non-minimal structures and they can
lead to the effective theories
with non-universal soft parameters\cite{SST}.
Third it was pointed out that higher order corrections generally
destroy the minimal form of the K\"ahler potential\cite{Rad-cor}.
Last the effects of supersymmetric grand unified theory (SUSY-GUT)
were little considered in the analysis of the running of
parameters although SUSY-GUT\cite{SUSY-GUT} has been hopeful
as a realistic theory.
We shall discuss the last point still more.
The unification dogma\cite{GUT} has a merit that the number of
independent parameters is reduced due to a large gauge symmetry.
Further SUSY $SU(5)$ GUT is supported by the LEP
experiments\cite{LEP} and predicts the long lifetime of nucleon
consistent with the present data\cite{Decay}.
Thus an analysis based on the MSSM encouraged by
the unification scenario seems to be hopeful.
In fact, many researches have been done under the assumption
that the soft SUSY breaking terms take a universal form
at the GUT scale $M_X$, but this assumption is also not
always realistic from the following reasons.
The non-minimal SUGRA can lead to the non-universal form
of soft SUSY breaking terms as described the above.
Even if we take a minimal SUGRA as a starting point,
the radiative correction from $M$ to $M_X$ changes
the universal form of SUSY breaking terms into non-universal one.
In some literatures, the renormalization effects
were discussed\cite{Ren-eff}, but we need to consider
effects on the gauge symmetry breaking further.
In Ref.~\cite{KMY2}, low-energy effective theory has
been derived from SUSY-GUT with non-universal soft SUSY
breaking terms by integrating out superheavy fields
and it is shown that new contributions to SUSY breaking terms
can appear at the tree level
after the breaking of unified gauge symmetry.
The analyses including the effects are started
recently\cite{Re-analyses}.
Now we should stress the importance of studying
the soft SUSY breaking terms.
The reason is that they can be a powerful probe to SUSY-GUT
and/or SUGRA since the weak scale SUSY spectrum
can directly reflect the physics at very high energies.
For example, we can check the GUT scenario experimentally
by measuring the gaugino masses\cite{Gaugino}.
Also, the scalar mass spectrum has certain ``sum rules''
specific to symmetry breaking patterns\cite{Sfermion}.
Therefore, the precision measurements of SUSY spectrum
are very important.
And it is a meaningful subject to obtain the low-energy theory
in more general framework and to grasp the peculiarities
concerning on the SUSY breaking terms in advance.
Various types of low-energy theories were derived
as will be explained in the next section.
However its low-energy theory has not been completely
investigated by taking SUGRA with general structure and
unified gauge symmetry as a starting point.
It has been only studied
in some specific cases\cite{HLW}\cite{Drees}\cite{KMY2}.
For example, it is shown that the universality of scalar masses
is preserved in the SUGRA whose K\"{a}hler potential
has $U(n)$ symmetry among the $n$ chiral fields\cite{HLW}.
In Ref.\cite{KMY2},
the scalar potential was derived starting from a unified theory
with a certain type of non-universal soft SUSY breaking terms.
Such non-universal soft terms arise if we take the flat limit
of the SUGRA where the K\"{a}hler potential is a certain type
of non-minimal one and the superpotential
is separate from the hidden sector to the visible one.
(We call this form of superpotential a ${\it hidden}$ ansatz.)
In this paper, we derive the low-energy effective theory from
non-minimal SUGRA with unified gauge symmetry.
The starting SUGRA has more general structure, i.e.,
the K\"{a}hler potential is non-minimal and
we do not impose the ${\it hidden}$ assumption
on the superpotential.
Then dangerous terms, which destabilize the gauge hierarchy,
generally appear at the tree level.
We discuss conditions that the hierarchy is preserved, and
take the flat limit and integrate out superheavy fields without
identifying $M_X$ with $M$.
We find various contributions to the SUSY breaking terms.
Our result reduces to that obtained in Ref.~\cite{HLW} in
the case with the minimal SUGRA.
Also it is shown that it reduces to that obtained
in Ref.\cite{KMY2} in the limit $M_X/M \to 0$
when we take a certain type of total K\"{a}hler potential.
The paper is organized as follows.
In section 2, we first review the low-energy effective
Lagrangians from SUGRA following the historical development.
We derive the low-energy effective scalar potential
starting from SUGRA with general total K\"{a}hler potential and
unified gauge symmetry in section 3.
In section 4, we discuss $D$-term contributions to
scalar masses and make clear the relation between our result
and that in Ref.\cite{KMY2}.
Section 5 is devoted to conclusions.
\section{Historical Background}
\subsection{Scalar Sector in SUGRA}
We begin by reviewing the scalar sector in SUGRA\cite{SUGRA}.
It is specified by two functions, the total K\"ahler
potential $G(\Phi, \Phi^*)$ and the gauge kinetic function
$f_{\alpha \beta}(\Phi)$ with $\alpha$, $\beta$
being indices of the adjoint representation of the gauge group.
The former is a sum of the K\"ahler potential $K$
and (the logarithm of) the superpotential $W_{SG}$ such as
\begin{eqnarray}
G(\Phi, \Phi^*)=K(\Phi, \Phi^*)
+M^{2}\ln |W_{SG}(\Phi) /M^{3}|^2.
\label{G}
\end{eqnarray}
We have denoted the chiral multiplets by $\Phi^{I}$
and their complex conjugate by $\Phi_{J}^*$.
The scalar potential is given by
\begin{eqnarray}
V= M^{2}e^{G/M^{2}} U
+\frac{1}{2} (Re f^{-1})_{\alpha \beta} D^{\alpha} D^{\beta},
\label{V}
\end{eqnarray}
where
\begin{eqnarray}
U &=& G^I (K^{-1})_I^J G_{J}-3M^{2},
\label{U}
\\
D^\alpha &=& G_I( T^\alpha z)^I
= (z^\dagger T^\alpha)_J G^J.
\label{D}
\end{eqnarray}
Here $G_{I}=\partial G/\partial z^I$,
$G^{J}=\partial G/\partial z_{J}^*$ etc, and
$T^\alpha$ are gauge transformation generators.
The $z^I$ is a scalar component of $\Phi^{I}$.
Here and hereafter both $G$ and $f_{\alpha \beta}$ are regarded
as functions of $z$ and $z^*$ as we take notice
of the scalar potential alone.
Also $(Re f^{-1})_{\alpha \beta}$ and $(K^{-1})_I^{J}$
are the inverse matrices of $Re f_{\alpha \beta}$ and $K_I^{J}$
respectively, and summation over $\alpha$,... and $I$,... is
understood. The last equality in Eq.~(\ref{D}) comes from the
gauge invariance of the total K\"ahler potential.
Let us next summarize our assumptions on the SUSY breaking.
The gravitino mass $m_{3/2}$ is given by
\begin{eqnarray}
m_{3/2} = \langle e^{K/2M^{2}} {W_{SG} \over M^2} \rangle,
\label{m}
\end{eqnarray}
where $\langle \cdots \rangle$ denotes
the vacuum expectation value (VEV) of the quantity.
We identify the gravitino mass with the weak scale.
The $F$-auxiliary fields of the chiral multiplets $\Phi^I$
are defined as
\begin{eqnarray}
F^I \equiv Me^{G/2M^{2}} (K^{-1})_J^{I} G^J.
\label{F}
\end{eqnarray}
We require those VEVs should satisfy
\begin{eqnarray}
\langle F^I \rangle \leq O(m_{3/2}M).
\label{<F>}
\end{eqnarray}
We can show that the VEVs of the $D$-auxiliary fields become
very small $\langle D^\alpha \rangle \leq O(m_{3/2}^2)$
as will be shown in Appendix A.
It follows from
Eqs.~(\ref{U}), (\ref{F}) and (\ref{<F>}) that
\begin{eqnarray}
\langle G_I \rangle, \ \langle G^J \rangle
\leq O(M)
\label{<G>}
\end{eqnarray}
and
\begin{eqnarray}
\langle U \rangle \leq O(M^{2}).
\label{<U>}
\end{eqnarray}
Note that we allow the non-zero
vacuum energy $\langle V \rangle$ of order $m_{3/2}^2 M^2$
at this level, which could be canceled by quantum corrections.
We also assume that derivatives of the K\"ahler potential
with respect to $z$ and $z^*$ are at most of order unity
(in the units where $M$ is taken to be unity), namely
\begin{eqnarray}
\langle K_{I_1 \cdots}^{J_1 \cdots} \rangle \leq O(1).
\label{KI1J1...}
\end{eqnarray}
This will be justified if the Planck scale physics plays
an essential role in the SUSY breaking.
We shall call the fields which induce to the SUSY breaking
$\lq$hidden fields', and denote those scalar components
and $F$-components as $\tilde{z}^i$
and $\tilde{F}^i$, respectively.
We require those VEVs should satisfy
\begin{eqnarray}
\langle \tilde{z}^i \rangle &=& O(M)
\label{<zi>}
\end{eqnarray}
and
\begin{eqnarray}
\langle \tilde{F}^i \rangle &=& O(m_{3/2} M).
\label{<Fi>}
\end{eqnarray}
We shall call the rest $\lq$observable fields' and denote
the scalar components as $z^\kappa$.
\subsection{Effective Theories from Minimal SUGRA}
The minimal SUGRA is defined as follows.
The K\"ahler potential $K$ has a canonical form as
\begin{eqnarray}
K = |z^{\kappa}|^2+|\tilde{z}^{i}|^2.
\label{Min-K}
\end{eqnarray}
We take the {\it hidden} ansatz for the superpotential as
\begin{eqnarray}
W_{SG} = W(z) + \tilde{W}(\tilde{z}).
\label{Hid-W}
\end{eqnarray}
The global SUSY theory with soft SUSY breaking terms
is derived by taking the flat limit,
i.e., $M \to \infty$ but $m_{3/2}$ kept finite.
The scalar potential is as follows\cite{Hidden},
\begin{eqnarray}
V &=& V_{SUSY} + V_{Soft},
\label{Min-V}\\
V_{SUSY} &=& |\frac{\partial \widehat{W}}{\partial z^\kappa}|^2
+ {1 \over 2}g_\alpha^2
(z_\kappa^* (T^\alpha)^\kappa_\lambda z^\lambda)^2 ,
\label{Min-VSUSY}\\
V_{Soft} &=& A \widehat{W}
+ B z^\kappa \frac{\partial \widehat{W}}{\partial z^\kappa}
+ {\it H.c.}
+ |B|^2 z_\kappa^* z^\kappa,
\label{Min-Vsoft}
\end{eqnarray}
where $\widehat{W}$ is defined as $\widehat{W}
\equiv \langle exp({K \over 2M^2})\rangle W$.
$V_{SUSY}$ stands for the supersymmetric part, while $V_{Soft}$
contains the soft SUSY breaking terms.
The $A$ and $B$ are the soft SUSY breaking parameters and
are written as
\begin{eqnarray}
A &=& {\langle \tilde{F}^i \rangle \langle K_i \rangle \over M^2}
- 3m_{3/2}^{\ast},
\label{A}\\
B &=& m_{3/2}^{\ast}.
\label{B}
\end{eqnarray}
This form of SUSY breaking terms is referred to as ``universal''.
The low-energy effective Lagrangian derived from the minimal SUGRA
with unified gauge symmetry also has a simple structure.
It was obtained by taking the flat limit and integrating out
superheavy fields simultaneously on the postulation that
the unification scale $M_X$ is identified with $M$\cite{HLW}.
The low-energy scalar potential takes the following form,
\begin{eqnarray}
V^{eff} &=& V_{SUSY}^{eff} + V_{Soft}^{eff},
\label{Veff}\\
V_{SUSY}^{eff} &=& |\frac{\partial \widehat{W}_{eff}}
{\partial z^k}|^2 + \frac{1}{2} g_a^2 (z_k^* (T^a)^k_l z^l)^2 ,
\label{VeffSUSY}
\\
V_{Soft}^{eff}
&=& A \widehat{W}_{\it eff}
+ B z^k \frac{\partial \widehat{W}_{\it eff}}
{\partial z^k } + {\it H.c.}
+ |B|^2 z_k^* z^k + \Delta V,
\label{VeffSoft}
\\
\Delta V &\equiv& -3A \widehat{W}_{\it eff}
+ A z^k \frac{\partial \widehat{W}_{\it eff}}
{\partial z^k } + {\it H.c.}
\label{DeltaV}
\end{eqnarray}
and it still has the same form as the original one
by a suitable redefinition of the $A$ and $B$ parameters
except the mass squared terms.
Here $z^k$ are the light scalar fields,\footnote{
They assumed that the supersymmetric masses of light fields from
the superpotential are zero.
It is straightforward to generalize their analysis into the case
that the light fields have non-zero but $O(m_{3/2})$ masses.}
$a$ is the index of generators of unbroken gauge group and
$\widehat{W}_{\it eff}$ is the superpotential $\widehat{W}$
with the extremum values for superheavy fields plugged in.
The scalar mass terms
are still universal with the same mass $B$.\footnote{
Throughout this subsection, it is assumed that the vacuum energy
$\langle V \rangle$ vanishes. In the presence of vacuum energy,
the value of scalar mass $|B|^2$ is replaced by
$|B|^2 + \langle V \rangle/M^2$.}
The universal structure of the low-energy Lagrangian led to
a number of strong conclusions, like the natural absence of
the flavor changing neutral currents \cite{FCNC}
or the radiative breaking scenario due to the heavy top
quark\cite{Rad-br}.
Due to these successes, the phenomenological analysis has
been made in popular based on the SUSY models with
the universal soft SUSY breaking terms\cite{MSSM}.
However, it becomes increasingly apparent that SUGRA
may not have the minimal form, and it is important to study
the consequences on the low-energy effective Lagrangian.
\subsection{Effective Theories from Non-minimal SUGRA}
The scalar potential is also obtained from the non-minimal SUGRA
with no superheavy fields and the result is given as\cite{S&W},
\begin{eqnarray}
V^{(non)} &=& V_{SUSY}^{(non)} + V_{Soft}^{(non)},
\label{Non-V}\\
V_{SUSY}^{(non)} &=&
\frac{\partial \widehat{\cal W^*}}{\partial z^*_\kappa}
\langle (K^{-1})_\kappa^\lambda \rangle
\frac{\partial \widehat{\cal W}}{\partial z^\lambda}
+ {1 \over 2}g_\alpha^2 (\langle K_\kappa^\lambda \rangle
z_\lambda^* (T^\alpha)^\kappa_\mu z^\mu)^2,
\label{Non-VSUSY}\\
V_{Soft}^{(non)}&=& A \widehat{\cal W}
+ B^\kappa(z)
\langle (K^{-1})_\kappa^\lambda \rangle
\frac{\partial \widehat{\cal W}}{\partial z^\lambda}
+ {\it H.c.}
\nonumber \\
&~&+ B^\kappa(z)
\langle (K^{-1})_\kappa^\lambda \rangle B_\lambda(z) + C(z, z^*) ,
\label{Non-VSoft}
\end{eqnarray}
where
\begin{eqnarray}
B^\kappa (z) &=& m_{3/2}^*
\langle K^\kappa_\lambda \rangle z^{\lambda}
- K^{\kappa}_j \langle \tilde{F}^j \rangle ,
\label{Bkappa}
\\
C(z, z^*) &=& -\langle \tilde{F}^i \rangle \delta^2 K_i^j
\langle \tilde{F}^*_j \rangle
\nonumber \\
&~& +\{{1 \over M^2}\langle \tilde{F}^i \rangle
\langle K_i^j \rangle \langle \tilde{F}^*_j \rangle
- 3|m_{3/2}|^2\} \delta^2 K
\nonumber \\
&~& + m_{3/2} \langle \tilde{F}^i \rangle \delta^2 K_i + {\it H.c.}
\nonumber \\
&~& - A\{m_{3/2}H(z) -\langle \tilde{F}^*_i \rangle H^i(z) \}
+ {\it H.c.}
\label{C}
\end{eqnarray}
in the case that we take the {\it hidden} ansatz.
Here $\widehat{\cal W}$ is defined as
\begin{eqnarray}
\widehat{\cal W} \equiv \widehat{W} + m_{3/2} H(z)
- \langle \tilde{F}^*_i \rangle H^i(z),
\label{calW}
\end{eqnarray}
where $H$ is the holomorphic part of $z^{\kappa}$ in $K$.
And $\delta^2 K$, $\delta^2 K_i$ and $\delta^2 K_i^j$ are
the quantities of order $m_{3/2}^2$, $m_{3/2}^2/M$ and
$m_{3/2}^2/M^2$ in $K$, $K_i$ and $K_i^j$, respectively.
Note that the SUSY breaking terms show a non-universal form.
As an excellent feature, there is a natural explanation
for the origin of $\mu$ parameter of order $m_{3/2}$
($\sim$ 1TeV)\cite{G&M}.
That is, the second and third terms in Eq.~(\ref{calW}) correspond
to $\mu$-term with a phenomenologically suitable order.
It is also known that the effective SUGRAs with non-minimal
structure are derived from 4-dimensional string models and
most of them lead to non-universal soft SUSY
breaking terms\cite{SST}.
When the {\it hidden} ansatz is taken off,
the following extra terms should be added,
\begin{eqnarray}
\frac{\partial \widehat{\cal W^*}}{\partial \tilde{z}^*_i}
\langle (K^{-1})_i^j \rangle
\frac{\partial \widehat{\cal W}}{\partial z^j}
+ \Delta C(z, z^*)
+ \langle \tilde{F}^i \rangle
\frac{\partial \widehat{\cal W}}{\partial \tilde{z}^i}
+ {\it H.c.},
\label{ExtraV}
\end{eqnarray}
where $\Delta C(z, z^*)$ is a bilinear polynomial of $z$ and $z^*$.
The magnitude of third term and its hermitian conjugate
can be of order $m_{3/2}^3 M$, and so a large mixing mass of
Higgs doublets can be introduced.
Hence we need to impose the condition
\begin{eqnarray}
\langle \tilde{F}^i \rangle
\frac{\partial \widehat{\cal W}}{\partial \tilde{z}^i}
= O(m_{3/2}^4)
\label{gh}
\end{eqnarray}
to guarantee the stability of weak scale.
The effective theories based on non-minimal SUGRA
with unified gauge symmetry also have been studied
in some literatures, but a complete analysis has not been
carried out yet.
For example, Hall {\it et} {\it al.} showed that the universality
of scalar masses is preserved
in the SUGRA whose K\"{a}hler potential has $U(n)$ symmetry
among the $n$ chiral fields\cite{HLW}.
Drees studied the low-energy theory based on SUGRA with
a non-canonical kinetic function
parametrized by one chiral field which triggers the SUSY
breaking\cite{Drees}.
As a recent development, the effective theory has been derived
from SUSY-GUT with a certain type of non-universal soft SUSY
breaking terms by integrating out superheavy fields\cite{KMY2}.
This starting SUSY-GUT can be derived from a certain type of
non-minimal SUGRA with unified gauge symmetry by imposing
the {\it hidden} ansatz and taking the flat limit first.
The low-energy effective scalar potential is obtained as follows,
\begin{eqnarray}
V^{eff(non)} &=& V_{SUSY}^{eff(non)} + V_{Soft}^{eff(non)},
\label{Non-Veff}
\\
V_{SUSY}^{eff(non)} &=&
|\frac{\partial \widehat{\cal W}_{eff}}{\partial z^k}|^2
+ {1 \over 2}g_a^2 (z_k^* (T^a)^k_l z^l)^2,
\label{Non-VeffSUSY} \\
V_{Soft}^{eff(non)}&=& A \widehat{\cal W}_{eff}
+ B^k(z)_{eff} \frac{\partial \widehat{\cal W}_{eff}}
{\partial z^k} + {\it H.c.}
\nonumber \\
&~& + B^k(z)_{eff} {B_k(z)}_{eff} + C(z, z^*)_{eff}
+ \Delta V^{(non)},
\label{Non-VeffSoft}
\end{eqnarray}
where
$\widehat{\cal W}_{\it eff}$,
$B^k(z)_{eff}$ and $C(z, z^*)_{eff}$ are
$\widehat{\cal W}$, $B^k(z)$ and $C(z, z^*)$
with the extremum values for superheavy fields plugged in,
and $\Delta V^{(non)}$ is a sum of extra contributions.
There exist new contributions specific to SUSY-GUTs with
non-universal soft SUSY breaking terms.
The appearance of $D$-term contribution to the scalar masses
is one example.\footnote{
Historically, it was demonstrated that the $D$-term contribution
occurs when the gauge symmetry is broken at an intermediate scale
due to the non-universal soft scalar masses in
Refs.~\cite{Hagelin} and its existence in a more general situation
was suggested in Ref.~\cite{Faraggi}.}
In the absence of Fayet-Iliopoulos $D$-term, the conditions
that sizable $D$-term contributions appear are as follows.
(1) SUSY-GUT has non-universal soft SUSY breaking terms.
(2) The rank of the gauge group is reduced by the gauge symmetry
breaking.
As the other feature, the gauge hierarchy achieved by a
fine-tuning in the superpotential would be violated, in general,
due to the non-universal SUSY breaking terms.
It is, however, shown that it is preserved for SUSY-GUT models
derived from the SUGRA with
the {\em hidden} ansatz and no light observable singlets.
It is also discussed some
phenomenological implications on the non-universal SUSY
breaking terms, including the utility of sfermion masses
as a probe of gauge symmetry breaking patterns and
the predictions of the radiative electroweak symmetry
breaking scenario and of no-scale type models.
As described in introduction,
it is important to study the low-energy theory in more general
framework of SUGRA because the SUSY spectrum can be a powerful
probe to the physics at higher energy scales.
The following subject has not been enough considered yet:
to obtain the low-energy theory directly from non-minimal
SUGRA with unified gauge symmetry in model-independent manner.
In the following sections, we carry it out paying attention to
the gauge hierarchy problem.
And we discuss extra contributions to the SUSY breaking terms and
the relation between our result and the previous one.
\section{Derivation of the Effective Lagrangian}
\subsection{Basic Assumptions}
\label{subsec:general-argument}
We have already explained general assumptions in the hidden
sector SUSY breaking scenario in subsection 2.1.
We shall first add basic assumptions although parts of them
would be repeated.
\begin{enumerate}
\item At the gravitational scale $M$,
the theory is described effectively as non-minimal
SUGRA with a certain unified gauge symmetry
whose K\"ahler potential and superpotential are
given as
\begin{eqnarray}
K &=& {K}(z, z^*; \tilde{z}, \tilde{z}^*)
\nonumber \\
&=& \tilde{K}(\tilde{z}, \tilde{z}^*)
+ \Lambda(z, z^*; \tilde{z}, \tilde{z}^*)
\nonumber \\
&~& + {H}(z ; \tilde{z}, \tilde{z}^*) + H.c.
\label{K}
\end{eqnarray}
and
\begin{eqnarray}
W_{SG} &=& W_{SG}(z, \tilde{z})
\nonumber \\
&=& \tilde{W}(\tilde{z}) + W(z, \tilde{z}),
\label{W} \\
W(z, \tilde{z}) &\equiv& {1 \over 2}m_{\kappa\lambda}(\tilde{z})
z^{\kappa}z^{\lambda} + {1 \over 3!}f_{\kappa\lambda\mu}
(\tilde{z})z^{\kappa}z^{\lambda}z^{\mu} + \cdots ,
\end{eqnarray}
respectively.
Here the dots stands for terms of higher orders in $z$.
The gauge group is not necessarily grand-unified into a
simple group.
The theory has no Fayet-Iliopoulos $U(1)$ $D$-term
for simplicity.\footnote{
The extension of the theory with Fayet-Iliopoulos $D$-term
is straightforward. We discuss it in Appendix B.}
\item The SUSY is spontaneously broken by the $F$-term
condensation in the hidden sector.
The Planck scale physics plays an essential role
in the SUSY breaking.\footnote{
Our discussion is also applicable to the case of SUSY breaking
by gaugino condensation if the freedoms are effectively
replaced by some scalar multiplets whose VEVs are of order $M$
as the models derived from SST.}
The hidden fields are gauge singlets
and they have the VEVs of $O(M)$.
The magnitude of ${W}_{SG}$ and $F$-component $\tilde{F}^i$ of
$\tilde{z}^i$ are $O(m_{3/2} M^2)$ and $O(m_{3/2} M)$, respectively.
\item The unified gauge symmetry is broken at a scale $M_X$.
Some observable scalar fields have the VEVs of $O(M_X)$.
\item All the particles can be classified as heavy
(with mass $O(M_X)$) or light (with mass $O(m_{3/2})$).
The light observable fields are gauge non-singlets\footnote{
The reason why we assume it is that there is a difficulty that
radiative corrections generally induce a large tadpole contribution
to Higgs masses in several models with a light singlet coupled to
Higgs doublets renormalizably in superpotential.}
and have fluctuations only of $O(m_{3/2})$.
\end{enumerate}
\subsection{Vacuum Solutions}
The scalar potential is given as
\begin{eqnarray}
V &=& V^{(F)} + V^{(D)},
\label{Vagain}
\\
V^{(F)} &\equiv& M^{2}exp(G/M^{2})(G^I (G^{-1})_I^J G_{J}-3M^{2})
\nonumber \\
&\equiv& M^{2}exp(G/M^{2}) U,
\label{V(F)}
\\
V^{(D)} &\equiv& \frac{1}{2} (Re f^{-1})_{\alpha \beta}
D^{\alpha} D^{\beta}.
\label{V(D)}
\end{eqnarray}
The index $I$, $J$,... run all scalar species, $i$, $j$,... run
the hidden fields and $\kappa$, $\lambda$,... run the observable
fields.
The $D^{\alpha}$'s are deformed as
\begin{eqnarray}
D^\alpha &=& K_\kappa (T^\alpha z)^\kappa
= (z^{\dagger}T^\alpha)_\kappa K^\kappa
\label{Dagain}
\end{eqnarray}
from the gauge invariance of superpotential.\footnote{
Note that the superpotential is not gauge invariant
under Fayet-Iliopoulos $U(1)$ transformation.}
The vacuum $\langle z^I \rangle$
and $\langle z^*_J \rangle$ are determined
by solving the stationary conditions $\partial V /
\partial z^I = 0$ and $\partial V / \partial z^*_J = 0$.
The conditions that the SUSY is not spontaneously broken
in the observable sector are simply expressed as
\begin{eqnarray}
\frac{\partial W}{\partial z^\kappa} &=& 0,
\label{SUSYF}\\
{D}^\alpha &=& 0.
\label{SUSYD}
\end{eqnarray}
We denote the solutions of the above conditions as
$z^{\kappa} = z_0^{\kappa}$.
We assume that our vacuum solution $\langle z^{\kappa} \rangle$ is
near to $z_0^{\kappa}$, i.e.,
$\langle z^{\kappa} \rangle = z_0^{\kappa}
+ O(m_{3/2})$.\footnote{
We can show that there exists at least such a vacuum solution
in the case that the scalar potential has no flat directions
in the SUSY limit.}
The supersymmetric fermion mass $\mu_{IJ}$ is given as
\begin{eqnarray}
\mu_{IJ} &=& \langle Me^{G/2M^2} ( G_{IJ} + {G_{I}G_{J} \over M^2}
- G_{I'}(G^{-1})^{I'}_{J'} G^{J'}_{IJ} ) \rangle.
\label{muIJ}
\end{eqnarray}
We take a basis of $z^I$ to diagonalize the SUSY
fermion mass matrix $\mu_{IJ}$.
Then we assume that the scalar fields are classified
either as ``heavy'' fields $z^K, z^L, \cdots$, ``light'' fields
$z^k, z^l, \cdots$, $z^i, z^j, \cdots$ such as $\mu_{KL}=O(M_X)$,
$\mu_{kl}=O(m_{3/2})$, $\mu_{ij}=O(m_{3/2})$ or Nambu--Goldstone
fields $z^A, z^B, \cdots$ (which will be discussed just below).
It is shown that the hidden fields belong to the light sector
in Appendix A.
The mass matrix of the gauge bosons $(M_V^2)^{\alpha\beta}$ is
given as
\begin{eqnarray}
(M_V^2)^{\alpha\beta} =
2 \langle (z^\dagger T^\beta)_\kappa K^\kappa_\lambda
(T^\alpha z)^\lambda \rangle,
\label{MV2}
\end{eqnarray}
up to the normalization due to the gauge coupling constants and
it can be diagonalized so that the gauge generators are
classified into ``heavy'' (those broken at $M_X$) $T^A, T^B,
\cdots$ and ``light'' (which remain unbroken above $m_{3/2}$)
$T^a, T^b, \cdots$. For the heavy generators, the fields
$\langle (T^A z)^\kappa \rangle$ correspond to the directions of
the Nambu--Goldstone fields in the field space, which span a vector
space with the same dimension as the number of heavy generators.
We can take a basis of the Nambu--Goldstone multiplets, $z^A, z^B,
\cdots$ so that
\begin{eqnarray}
\sqrt{2} \langle (T^A z)^B \rangle = M_V^{AB} .
\label{NG}
\end{eqnarray}
Here the Nambu--Goldstone fields are taken to be orthogonal to
the heavy and light fields such as $\langle (T^A z)^K
\rangle = 0$ and
$\langle (T^A z)^k \rangle = 0$.
To be more precise, either real or imaginary parts of
the $z^A$'s are the true Nambu--Goldstone bosons which are absorbed
into the gauge bosons, and the other parts acquire the same mass
of order $M_X$ as that of the gauge bosons from the $D$-term
$V^{(D)}$ in the SUSY limit.
Hence the Nambu--Goldstone multiplets belong to the heavy sector.
Let us give the procedure to obtain the low-energy
effective theory.
\begin{enumerate}
\item We calculate the VEVs of the derivatives of the potential
and we write down the potential as
\begin{eqnarray}
V=\frac{1}{2} \langle V_{IJ} \rangle
\Delta z^I \Delta z^J +\cdots ,
\end{eqnarray}
where the scalar fields $z^I$'s are expanded
as $z^I = \langle z^I \rangle + \Delta z^I$
around the vacuum $\langle z^I \rangle$.
\item When there exists a mass mixing between the heavy and
light sectors, we need to diagonalize them to identify the
light and heavy fields correctly.
\item We solve the stationary conditions of the potential
for the heavy scalar fields while keeping the light scalar fields
arbitrary and then integrate out the heavy fields by inserting
the solutions of the stationary conditions into the potential.
We take the flat limit simultaneously.
\end{enumerate}
\subsection{Derivatives of $K$ and $W$}
It is convenient to
write both the K\"ahler potential $K$ and the superpotential
$W_{SG}$ in terms of the variations $\Delta z^I$ and
$\Delta z^*_J$ as follows,
\begin{eqnarray}
K &=& \langle K \rangle + \langle K_I \rangle \Delta z^I
+ \langle K^J \rangle
\Delta z^*_J
\nonumber \\
&~& + \langle K_I^J \rangle \Delta z^I \Delta z^*_J
\nonumber \\
&~& + {1 \over 2}\langle K_{IJ} \rangle \Delta z^I \Delta z^J
+ {1 \over 2}\langle K^{IJ} \rangle \Delta z^*_I \Delta z^*_J
\nonumber \\
&~&+ \cdots
\label{expK}
\end{eqnarray}
and
\begin{eqnarray}
W_{SG} &=& \langle W_{SG} \rangle + \langle W_{SG I} \rangle
\Delta z^I
\nonumber \\
&~&+ {1 \over 2}\langle W_{SG IJ} \rangle \Delta z^I \Delta z^J
+ {1 \over 3!}\langle W_{SG IJJ'} \rangle \Delta z^I \Delta z^J
\Delta z^{J'}
\nonumber \\
&~&+ \cdots ,
\label{expW}
\end{eqnarray}
where the ellipses represent higher order terms in $\Delta z$.
By using the expansions (\ref{expK}) and (\ref{expW}),
we find the following estimations
\begin{eqnarray}
&~& \langle G_i \rangle = O(M), \ \langle G_K \rangle \leq O(M_X),
\nonumber \\
&~& \langle G_A \rangle \leq O(m_{3/2}^2/M_X),
\ \langle G_k \rangle =0,
\label{GI}\\
&~& \langle G_i^j \rangle \leq O(1), \
\langle G_{\kappa}^{\lambda} \rangle \leq O(1), \
\langle G_{K}^j \rangle \leq O(M_X/M),
\nonumber \\
&~& \langle G_{A}^j \rangle \leq O(M_X/M),
\ \langle G_{k}^j \rangle = 0,
\label{GI^J}\\
&~& \langle G_{ij} \rangle \leq O(1), \
\langle G_{KL} \rangle \leq O(M_{KL}/m_{3/2}), \
\langle G_{Kj} \rangle \leq O(M_X/M),
\nonumber \\
&~& \langle G_{Aj} \rangle \leq O(M_X/M),
\ \langle G_{kj} \rangle = 0,
\nonumber \\
&~& \langle G_{\kappa B} \rangle \leq O(1),
\ \langle G_{\kappa l} \rangle
\leq O(1),
\label{GIJ}
\end{eqnarray}
where $M_{KL}$ is the SUSY fermion mass coming
from the superpotential.
Here we used the assumption that our vacuum solution is near to
that in the SUSY limit and a perturbative argument
to derive the second relation in (\ref{GI}).
And we used the relations (\ref{NG}) and
$\langle D^{\alpha} \rangle
\leq O(m_{3/2}^2)$ to derive the third relation in (\ref{GI}).
By using the equality from the gauge invariance (\ref{G-inv2}),
we derive the following relations,
\begin{eqnarray}
&~&\langle G_{Akl} \rangle \leq O(1/M_X),
\ \langle G_{ABl} \rangle \leq O(1/M_X),
\nonumber \\
&~& \langle G_{ABC} \rangle \leq O(1/M_X),
\ \langle G_{Akj} \rangle \leq O(1/M),
\nonumber \\
&~& \langle G_{ABj} \rangle \leq O(1/M),
\ \langle G_{Aij} \rangle \leq O(1/M)
\label{GAIJ}
\end{eqnarray}
or
\begin{eqnarray}
&~&\langle {W_{SG}}_{Akl} \rangle \leq O(m_{3/2}/M_X),\
\langle {W_{SG}}_{ABl} \rangle \leq O(m_{3/2}/M_X),
\nonumber \\
&~& \langle {W_{SG}}_{ABC} \rangle \leq O(m_{3/2}/M_X),\
\langle {W_{SG}}_{Akj} \rangle \leq O(m_{3/2}/M),
\nonumber \\
&~& \langle {W_{SG}}_{ABj} \rangle \leq O(m_{3/2}/M), \
\langle {W_{SG}}_{Aij} \rangle \leq O(m_{3/2}/M).
\label{WAIJ}
\end{eqnarray}
\subsection{Stability of Gauge Hierarchy}
The mass squared matrices of the scalar fields are simply given
by the VEVs of the second derivatives of the potential. From
Eqs. (\ref{Vagain})--(\ref{Dagain}), we get the relations,
\begin{eqnarray}
\lefteqn{V_{I}^{J}=
\frac{\partial ^2 V}{\partial \phi^I \partial \phi^*_J}=}
\nonumber \\
& & M^{2}(e^{G/M^{2}})_{I}^{J} U
+M^{2}(e^{G/M^{2}})_I U^{J}
+M^{2}(e^{G/M^{2}})^J U_I
+M^{2}e^{G/M^{2}} U_{I}^{J}
\nonumber \\
& & +\frac{1}{2} (Re f^{-1})_{\alpha \beta, I}^{J}
D^\alpha D^\beta +(Re f^{-1})_{\alpha \beta, I}
D^\alpha (D^\beta)^J +(Re f^{-1})_{\alpha \beta}^J
D^\alpha (D^\beta)_I
\nonumber \\
& & +(Re f^{-1})_{\alpha \beta }
D^\alpha (D^\beta)_{I}^{J}+(Re f^{-1})_{\alpha \beta}
(D^\alpha)_I (D^\beta)^J
\label{VI^J}
\end{eqnarray}
and
\begin{eqnarray}
\lefteqn{V_{IJ}=
\frac{\partial ^2 V}{\partial \phi^I \partial \phi^J}=}
\nonumber \\
& & M^{2}(e^{G/M^{2}})_{IJ} U
+M^{2}(e^{G/M^{2}})_I U_J
+M^{2}(e^{G/M^{2}})_J U_I
+M^{2}e^{G/M^{2}} U_{IJ}
\nonumber \\
& & +\frac{1}{2} (Re f^{-1})_{\alpha \beta, IJ}
D^\alpha D^\beta+(Re f^{-1})_{\alpha \beta, I}
D^\alpha (D^\beta)_J+(Re f^{-1})_{\alpha \beta, J}
D^\alpha (D^\beta)_I
\nonumber \\
& & +(Re f^{-1})_{\alpha \beta }
D^\alpha (D^\beta)_{IJ}+(Re f^{-1})_{\alpha \beta}
(D^\alpha)_I (D^\beta)_J.
\label{VIJ}
\end{eqnarray}
By using the relations (\ref{GI})--(\ref{WAIJ}),
the VEVs of $V_{I}^{J}$ and $V_{IJ}$ are estimated as
\footnote{For simplicity, hereafter we consider only
the case that the equality holds.}
\begin{eqnarray}
&~&\langle {V^{(F)}}_{K}^L \rangle = O(M_X^2),
\ \langle {V^{(F)}}_{A}^B \rangle = O(m_{3/2}^2),\
\langle {V^{(F)}}_{k}^l \rangle = O(m_{3/2}^2),
\nonumber \\
&~&\langle {V^{(F)}}_{i}^j \rangle = O(m_{3/2}^2),\
\langle {V^{(F)}}_{K}^B \rangle = O(m_{3/2} M_X),\
\langle {V^{(F)}}_{K}^l \rangle = O(m_{3/2} M_X),
\nonumber \\
&~&\langle {V^{(F)}}_{K}^j \rangle = O(m_{3/2}M_X^2/M),\
\langle {V^{(F)}}_{A}^l \rangle = O(m_{3/2}^2),\
\nonumber \\
&~&\langle {V^{(F)}}_{A}^j \rangle = O(m_{3/2}^2 M_X/M),\
\langle {V^{(F)}}_{k}^j \rangle = 0
\label{<VI^J>F}
\end{eqnarray}
and
\begin{eqnarray}
&~&\langle V^{(F)}_{KL} \rangle = O(m_{3/2} M),
\ \langle V^{(F)}_{AB} \rangle = O(m_{3/2}^2),\
\langle V^{(F)}_{kl} \rangle = O(m_{3/2} M),
\nonumber \\
&~&\langle V^{(F)}_{ij} \rangle = O(m_{3/2} M),\
\langle V^{(F)}_{KB} \rangle = O(m_{3/2} M),\
\langle V^{(F)}_{Kl} \rangle = O(m_{3/2} M),
\nonumber \\
&~&\langle V^{(F)}_{Kj} \rangle = O(m_{3/2} M),\
\langle V^{(F)}_{Al} \rangle = O(m_{3/2}^2),\
\langle V^{(F)}_{Aj} \rangle = O(m_{3/2}^2),
\nonumber \\
&~&\langle V^{(F)}_{kj} \rangle = 0,
\label{<VIJ>F}
\end{eqnarray}
respectively.
The quantities of order $m_{3/2} M$ in
$\langle V^{(F)}_{IJ} \rangle$
originates in the term $\langle Me^{G/2M^{2}}G_{IJJ'} \rangle
\langle F^{J'} \rangle$.
If $\langle V_{IJ} \rangle$'s are $O(m_{3/2} M)$ for the light
fields $z^I$, the masses of light fields can get intermediate
values after the diagonalization of mass matrix.
The masses of those fermionic partners stay at the weak scale.
The weak scale can be destabilized in the presence of the weak
Higgs doublets with intermediate masses.
This is so called $\lq$gauge hierarchy problem'.
Only when $\langle Me^{G/2M^{2}}G_{IJJ'} \rangle
\langle F^{J'} \rangle$'s meet
some requirements, the hierarchy survives.
In this paper, we require the following conditions,
\begin{eqnarray}
\langle Me^{G/2M^{2}}G_{IJJ'} \rangle \langle {F}^{J'} \rangle \leq
O(m_{3/2}^2)
\label{gh1}
\end{eqnarray}
for the light fields $z^I$ and $z^J$,
\begin{eqnarray}
&~& \langle Me^{G/2M^{2}}G_{KlJ'} \rangle
\langle {F}^{J'} \rangle \leq
O(m_{3/2} M_X),
\label{gh3} \\
&~& \langle Me^{G/2M^{2}}G_{KjJ'} \rangle
\langle {F}^{J'} \rangle \leq
O(m_{3/2} M_X^2/M),
\label{gh4} \\
&~& \langle Me^{G/2M^{2}}G_{IJJ'} \rangle
\langle {F}^{J'} \rangle \leq
O(M_X^2)
\label{gh2}
\end{eqnarray}
for the heavy fields $z^I$ and $z^J$, and
\begin{eqnarray}
\langle Me^{G/2M^{2}}G_{AjJ'} \rangle
\langle {F}^{J'} \rangle \leq
O(m_{3/2}^2 M_X/M).
\label{gh5}
\end{eqnarray}
The conditions (\ref{gh1})--(\ref{gh5}) correspond to
the statement that the magnitudes of
$\langle Me^{G/2M^{2}}G_{IJJ'} \rangle \langle F^{J'} \rangle$
are equal to or smaller than the rest terms.
The {\it hidden} ansatz trivially satisfies the above conditions.
The gauge hierarchy problem has been discussed on the postulation
that $M_X$ is identified with $M$ in Ref.~\cite{JKY}.
The contributions from the $D$-term are naively estimated
as follows,
\begin{eqnarray}
&~&\langle {V^{(D)}}_{\kappa}^{\lambda} \rangle = O(M_X^2),
\ \langle V^{(D)}_{\kappa\lambda} \rangle = O(M_X^2),
\nonumber \\
&~&\langle {V^{(D)}}_{K}^j \rangle,
\ \langle {V^{(D)}}_{A}^j \rangle = O(M_X^3/M),
\ \langle V^{(D)}_{Kj} \rangle,
\ \langle V^{(D)}_{Aj} \rangle = O(M_X^3/M),
\nonumber \\
&~&\langle {V^{(D)}}_{i}^j \rangle,
\ \langle V^{(D)}_{ij} \rangle = O(M_X^4/M^2),
\nonumber \\
&~&\langle {V^{(D)}}_{k}^j \rangle,
\ \langle V^{(D)}_{kj} \rangle = 0,
\label{<VIJ>D}
\end{eqnarray}
where we used the relation $\langle (D^A)_I \rangle
= \langle (z^{\dagger} T^A)_{\kappa} K_I^{\kappa} \rangle
= O(M_X) K^A_I$.
We require that the light fields defined by using $\mu_{IJ}$
get no heavy masses from the $D$-term.
For simplicity, we impose the conditions such as
\begin{eqnarray}
\langle {V^{(D)}}_I^J \rangle \leq \langle {V^{(F)}}_I^J \rangle
\label{<VD><<VF>1}
\end{eqnarray}
and
\begin{eqnarray}
\langle V^{(D)}_{IJ} \rangle \leq \langle V^{(F)}_{IJ} \rangle
\label{<VD><<VF>2}
\end{eqnarray}
for the light fields $z^I$.
They yield to the following relations
\begin{eqnarray}
\langle K_{A}^{k} \rangle, \ \langle K_{Ak} \rangle
= O({m_{3/2}^2 \over M_X^2})
\label{<KAk>}
\end{eqnarray}
and
\begin{eqnarray}
\langle K_{A}^{i} \rangle, \ \langle K_{Ai} \rangle
= O({m_{3/2}^2 \over M_XM}).
\label{<KAi>}
\end{eqnarray}
The analysis could be made based on weaker requirements than
(\ref{gh1})--(\ref{gh5}), (\ref{<VD><<VF>1}) and
(\ref{<VD><<VF>2}), but we will not discuss it further to avoid
a complication and a subtlety in this paper.
\subsection{Diagonalization of Mass Matrix}
The mass term is written as
\begin{eqnarray}
V^{mass}=\frac{1}{2} \langle V_{\hat{I}\hat{J}} \rangle
{\Delta z}^{\hat{I}}{\Delta z}^{\hat{J}} ,
\label{Vmass}
\end{eqnarray}
where ${\Delta z}^{\hat{I}}$$=({\Delta z}^K, {\Delta z}^{\bar{L}}
\equiv {\Delta z}^*_L;
{\Delta z}^A, {\Delta z}^{\bar{B}} \equiv {\Delta z}^*_B;$
$\Delta \tilde{z}^i,$ $\Delta \tilde{z}^{\bar{j}}$ $\equiv$
$\Delta \tilde{z}^*_j;$ ${\Delta z}^k,$ ${\Delta z}^{\bar{l}}$
$\equiv$ ${\Delta z}^*_l)$. From the discussion in the previous
subsection, the orders of $\langle V_{\hat{I}\hat{J}} \rangle$
are estimated as
\begin{eqnarray}
\langle V_{\hat{I}\hat{J}} \rangle = O
\left(
\begin{array}{ccc}
M_X^2 & M_X^2 & m_{3/2} M_X \\
M_X^2 & M_X^2 & m_{3/2}^2 \\
m_{3/2} M_X & m_{3/2}^2 & m_{3/2}^2
\end{array}
\right)
\label{<VIJ>N}
\end{eqnarray}
for gauge non-singlet fields $({\Delta z}^{\hat{K}};
{\Delta z}^{\hat{A}}; {\Delta z}^{\hat{k}})$
and
\begin{eqnarray}
\langle V_{\hat{I}\hat{J}} \rangle = O
\left(
\begin{array}{ccc}
M_X^2 & M_X^2 & m_{3/2} M_X^2/M \\
M_X^2 & M_X^2 & m_{3/2}^2 M_X/M \\
m_{3/2} M_X^2/M & m_{3/2}^2 M_X/M & m_{3/2}^2
\end{array}
\right)
\label{<VIJ>S}
\end{eqnarray}
for gauge singlet fields $({\Delta z}^{\hat{K}};
{\Delta z}^{\hat{A}}; \Delta \tilde{z}^{\hat{i}})$.
As the matrix $\langle V_{\hat{I}\hat{J}} \rangle$ is hermitian,
it can be diagonalized by the use of a certain unitary matrix
$U^{\hat{I}}_{\hat{J}}$.
The mass eigenstate $\phi^{\hat{I}}$ is related to
${\Delta z}^{\hat{I}}$ as $\phi^{\hat{I}} = U^{\hat{I}}_{\hat{J}}
{\Delta z}^{\hat{J}}$.
We denote the heavy fields with mass $O(M_X)$ as
$\phi^{\hat{\cal H}}$ and the light fields with mass $O(m_{3/2})$
as $\phi^{\hat{\cal L}}$ where ${\cal H} = (K, A)$ and ${\cal L}
= (i, k)$. Next we would like to integrate out the heavy fields
$\phi^{\hat{\cal H}}$.
For this purpose, it is convenient to choose the variables
\begin{eqnarray}
\Delta \hat{z}^{\hat{\cal H}} &=&
(U^{\hat{\cal H}}_{\hat{\cal H'}})^{-1} \phi^{\hat{\cal H'}},
\label{hatzH}
\\
\Delta \hat{z}^{\hat{\cal L}} &=&
(U^{\hat{\cal L}}_{\hat{\cal L'}})^{-1} \phi^{\hat{\cal L'}}
\label{hatzL}
\end{eqnarray}
or
\begin{eqnarray}
\Delta \hat{z}^{\hat{I}} &=& \hat{U}^{\hat{I}}_{\hat{J}}
{\Delta z}^{\hat{J}},
\label{hatz}
\\
\hat{U}^{\hat{I}}_{\hat{J}} &\equiv&
\left(
\begin{array}{cc}
I & (U^{\hat{\cal H}}_{\hat{\cal H'}})^{-1}
U^{\hat{\cal H'}}_{\hat{\cal L}}\\
(U^{\hat{\cal L}}_{\hat{\cal L'}})^{-1}
U^{\hat{\cal L'}}_{\hat{\cal H}}
& I
\end{array}
\right).
\label{hatU}
\end{eqnarray}
Here we used the fact that $det U^{\hat{\cal H}}_{\hat{\cal H'}} =
1 + O(m_{3/2}^2/M_X^2)$ and $det U^{\hat{\cal L}}_{\hat{\cal L'}} =
1 + O(m_{3/2}^2/M_X^2)$ and neglected the higher order terms.
The orders of off-diagonal elements of
$\hat{U}^{\hat{I}}_{\hat{J}}$ are estimated as
\begin{eqnarray}
&~&\hat{U}^{\hat{K}}_{\hat{l}}
= O({m_{3/2} \over M_X}),~
\hat{U}^{\hat{A}}_{\hat{l}}
= O({m_{3/2}^2 \over M_X^2}),
\label{off1}
\\
&~&\hat{U}^{\hat{K}}_{\hat{j}}
= O({m_{3/2} \over M}),~
\hat{U}^{\hat{A}}_{\hat{j}}
= O({m_{3/2}^2 \over M M_X}).
\label{off2}
\end{eqnarray}
\subsection{Calculation of the Effective Theory}
The rest in the procedure are as follows,\\
1. We write down the scalar potential by using new variables
$\Delta \hat{z}^{\hat{I}}$.\\
2. We take the flat limit and integrate out the heavy fields
by inserting the solutions of the stationary conditions into
the full potential.
We can write down the K\"ahler potential $K$, the superpotential
$W_{SG}$ and the $D$-auxiliary fields $D^\alpha$ in terms of
the variations $\Delta \hat{z}^{\hat{I}}$ as follows,
\begin{eqnarray}
K &=&
\hat{K}(\Delta \hat{z})
\nonumber \\
&=&\langle \hat{K} \rangle + \langle \hat{K}_{\hat{I}} \rangle
\Delta \hat{z}^{\hat{I}} + {1 \over 2}\langle
\hat{K}_{\hat{I}\hat{J}} \rangle
\Delta \hat{z}^{\hat{I}} \Delta \hat{z}^{\hat{J}}
+ \cdots ,
\label{hatK}\\
W_{SG} &=&
\hat{W}(\Delta \hat{z})
\nonumber \\
&=&\langle \hat{W} \rangle + \langle \hat{W}_{\hat{I}} \rangle
\Delta \hat{z}^{\hat{I}} + {1 \over 2}\langle
\hat{W}_{\hat{I}\hat{J}} \rangle
\Delta \hat{z}^{\hat{I}} \Delta \hat{z}^{\hat{J}}
\nonumber \\
&~& + {1 \over 3!}\langle \hat{W}_{\hat{I}\hat{J}\hat{J'}}
\rangle \Delta \hat{z}^{\hat{I}} \Delta \hat{z}^{\hat{J}}
\Delta \hat{z}^{\hat{J'}}
+ \cdots
\label{hatW}
\end{eqnarray}
and
\begin{eqnarray}
D^{\alpha} &=& \hat{D}^{\alpha}(\Delta \hat{z})
\nonumber \\
&=& (\hat{K}_\lambda + \hat{K}_{\hat{I}}
\Delta \hat{U}^{\hat{I}}_{\lambda}) (T^A)_\kappa^\lambda
(\langle z^{\kappa} \rangle +
(\hat{U}^{-1})^{\kappa}_{\hat{J}}\Delta \hat{z}^{\hat{J}}) ,
\label{hatD}
\end{eqnarray}
where the ellipses represent terms of higher orders and
$\hat{U}^{\hat{I}}_{\hat{J}} =
\delta^{\hat{I}}_{\hat{J}} + \Delta \hat{U}^{\hat{I}}_{\hat{J}}$.
For a later convenience, we deform $V^{(F)}$ as follows,
\begin{eqnarray}
V^{(F)} &=&
exp(\hat{K}/M^{2})\biggl(\widehat{{\cal G}}_{\bar{\kappa}}
(\hat{K}^{-1})^{{\bar{\kappa}}{\lambda}}
\widehat{{\cal G}}_{\lambda}
\nonumber \\
&~&~~~~~~ + \hat{\cal G}_{\bar{i}} (\widehat{K}^{-1})^{{\bar{i}}j}
\hat{\cal G}_{j}
-3{|\hat{W}|^2 \over M^{2}}\biggr) + \Delta V^{(F)},
\label{V(F)again}
\end{eqnarray}
where
\begin{eqnarray}
\widehat{{\cal G}}_{\bar{\kappa}} &\equiv&
\hat{\cal G}_{\bar{\kappa}} + \hat{\cal G}_{\bar{i}}
(\hat{K}^{-1})^{{\bar{i}}\mu} (\hat{K})_{{\mu}{\bar{\kappa}}},
\label{widehatg*}
\\
\widehat{\cal G}_{\lambda} &\equiv&
\hat{\cal G}_{\lambda} + (\hat{K})_{\lambda\bar{\nu}}
(\hat{K}^{-1})^{{\bar{\nu}}j} \hat{\cal G}_j,
\label{widehatg}
\\
\hat{\cal G}_{\bar{I}} &\equiv& \hat{W}^*_{\bar{I}}
+ {\hat{K}_{\bar{I}} \over M^2}\hat{W}^* ,
\label{hatg*}
\\
\hat{\cal G}_{I} &\equiv& \hat{W}_{I}
+ {\hat{K}_{I} \over M^2}\hat{W}
\label{hatg}
\end{eqnarray}
and
\begin{eqnarray}
(\widehat{K}^{-1})^{{\bar{i}}j} &\equiv&
(\hat{K}^{-1})^{{\bar{i}}j}
- (\hat{K}^{-1})^{{\bar{i}}\mu} (\hat{K})_{\mu\bar{\nu}}
(\hat{K}^{-1})^{{\bar{\nu}}j},
\label{widehatK}
\\
\Delta V^{(F)} &\equiv&
exp(\hat{K}/M^{2})\biggl(\hat{\cal G}_{\hat{I}}
\Delta(\hat{U})^{\hat{I}}_{\bar{I}}
(\hat{K}^{-1})^{{\bar{I}}J} \hat{\cal G}_J
+\hat{\cal G}_{\bar{I}}(\hat{K}^{-1})^{{\bar{I}}J}
\hat{\cal G}_{\hat{J}} \Delta(\hat{U})^{\hat{J}}_J
\nonumber \\
&~&~~~~ +\hat{\cal G}_{\bar{I}}[(\hat{K}^{-1})^{{\hat{J}}J}
\Delta(\hat{U}^{-1})_{\hat{J}}^{\bar{I}}
+ (\hat{K}^{-1})^{{\hat{I}}{\bar{I}}}
\Delta(\hat{U}^{-1})^J_{\hat{I}}] \hat{\cal G}_{J} \biggr)
\nonumber \\
&~&~~~~ + O((\Delta U^{(-1)})^2).
\label{DeltaV(F)}
\end{eqnarray}
We should not confuse $(\widehat{K}^{-1})^{{\bar{i}}j}$,
$\widehat{{\cal G}}_{\bar{\kappa}}$
and $\widehat{\cal G}_{\lambda}$ with
$(\hat{K}^{-1})^{{\bar{i}}j}$, $\hat{{\cal G}}_{\bar{\kappa}}$
and $\hat{\cal G}_{\lambda}$, respectively.
(Notice that the difference of the size of hat.)
Here $(\hat{K})_{\mu{\bar{\nu}}}$ is
the inverse matrix of $(\hat{K}^{-1})^{\mu{\bar{\nu}}}$.
We expand $\Delta \hat{z}^{\hat{I}}$ in powers of $m_{3/2}$ such as
\begin{eqnarray}
\Delta \hat{z}^{\hat{I}} &=&
\delta \hat{z}^{\hat{I}} + \delta^2 \hat{z}^{\hat{I}}
+ \cdots,
\label{exphatz}
\end{eqnarray}
with $\delta^n \hat{z}^{\hat{I}} = O(m_{3/2}^n / M_X^{n-1})$.
We assume $\Delta \hat{z}^{\hat{k}} =
O(m_{3/2})$ for the light fields, {\it e.g.},\/
$\delta^2 \hat{z}^{\hat{k}} = \delta^3 \hat{z}^{\hat{k}}
= \cdots =0$.
In the same way, we expand the $\widehat{\cal G}_{\hat{\lambda}}$,
$\hat{\cal G}_{\hat{j}}$ and $\hat{D}^\alpha$
in powers of $m_{3/2}$ such as
\begin{eqnarray}
\widehat{\cal G}_{\hat{\lambda}} &=&
\delta \widehat{\cal G}_{\hat{\lambda}}
+ \delta^2 \widehat{\cal G}_{\hat{\lambda}} + \cdots,
\label{expwidehatg}\\
\hat{\cal G}_{\hat{j}} &=&
\delta \hat{\cal G}_{\hat{j}} + \delta^2 \hat{\cal G}_{\hat{j}}
+ \cdots \label{exphatg}
\end{eqnarray}
and
\begin{eqnarray}
\hat{D}^\alpha &=& \delta
\hat{D}^\alpha + \delta^2 \hat{D}^\alpha + \cdots .
\label{exphatD}
\end{eqnarray}
Those orders are given as $\delta^n \widehat{\cal G}_{\hat{\lambda}}
= O(m_{3/2}^n / M_X^{n-2})$,
$\delta^n \hat{\cal G}_{\hat{j}} = O(m_{3/2}^n / M^{n-2})$
and $\delta^n \hat{D}^\alpha$$=$$O(m_{3/2}^n / M_X^{n-2})$
up to the factor $O((M_X/M)^n)$.
The following relations are derived from the expansions of
$\widehat{\cal G}_\lambda$ and $\hat{\cal G}_{j}$
\begin{eqnarray}
\delta \widehat{\cal G}_K &=& \langle \hat{W}_K \rangle +
\langle \hat{W}_{KL} \rangle \delta \hat{z}^L
\nonumber \\
&~&+ {\langle \hat{W} \rangle \over M^2}\langle \hat{K}_K \rangle
+ \langle (\hat{K})_{K\bar{\nu}} \rangle \langle
(\hat{K}^{-1})^{{\bar{\nu}}j} \rangle \delta \hat{\cal G}_j,
\label{gK1}
\\
\delta^2 \widehat{\cal G}_K &=&
\langle \hat{W}_{KL} \rangle \delta^2 \hat{z}^L
+ \langle \hat{W}_{K{\cal L}} \rangle \delta \hat{z}^{\cal L}
+ \langle \hat{W}_{KA} \rangle \delta \hat{z}^A
\nonumber \\
&~& + {1 \over 2}\langle \hat{W}_{K\lambda\mu} \rangle
\delta \hat{z}^{\lambda} \delta \hat{z}^{\mu}
\nonumber \\
&~& + {1 \over M^2}\biggl({1 \over 2}\langle \hat{W}_{LM} \rangle
\delta \hat{z}^L \delta \hat{z}^M \langle \hat{K}_K \rangle
+ \langle \hat{W} \rangle \langle \hat{K}_{K\hat{J}} \rangle
\delta \hat{z}^{\hat{J}}\biggr)
\nonumber \\
&~& + \langle (\hat{K})_{K\bar{\nu}} \rangle
\langle (\hat{K}^{-1})^{{\bar{\nu}}j} \rangle
\delta^2 \hat{\cal G}_j
+ \delta ((\hat{K})_{K\bar{\nu}} (\hat{K}^{-1})^{{\bar{\nu}}j})
\delta \hat{\cal G}_j,
\label{gK2}
\\
\delta \widehat{\cal G}_A &=& \langle (\hat{K})_{A\bar{\nu}}
\rangle \langle (\hat{K}^{-1})^{{\bar{\nu}}j} \rangle
\delta \hat{\cal G}_j,
\label{gA1}
\\
\delta^2 \widehat{\cal G}_A &=&
\langle \hat{W}_{AI} \rangle \delta \hat{z}^{I}
+ {1 \over 2}\langle \hat{W}_{AKI} \rangle
\delta \hat{z}^{K} \delta \hat{z}^{I}
+ {\langle \hat{W} \rangle \over M^2}
\langle \hat{K}_{A\hat{J}} \rangle \delta \hat{z}^{\hat{J}}
\nonumber \\
&~& + \langle (\hat{K})_{A\bar{\nu}} \rangle
\langle (\hat{K}^{-1})^{{\bar{\nu}}j} \rangle
\delta^2 \hat{\cal G}_j
+ \delta ((\hat{K})_{A\bar{\nu}} (\hat{K}^{-1})^{{\bar{\nu}}j})
\delta \hat{\cal G}_j
\label{gA2}
\end{eqnarray}
and
\begin{eqnarray}
\delta {\hat{\cal G}_j} &=& \langle \hat{W}_j \rangle
+{\langle \hat{W} \rangle \over M^2} \langle \hat{K}_{j} \rangle,
\label{gj1}
\\
\delta^2 {\hat{\cal G}_j} &=&
\langle \hat{W}_{jI} \rangle \delta \hat{z}^{I}
+ {1 \over 2}\langle \hat{W}_{jIJ} \rangle
\delta \hat{z}^{I} \delta \hat{z}^{J}
\nonumber \\
&~& + {1 \over M^2}\biggl({1 \over 2}\langle \hat{W}_{LM} \rangle
\delta \hat{z}^L \delta \hat{z}^M \langle \hat{K}_j \rangle
+ \langle \hat{W} \rangle
\langle \hat{K}_{j\hat{J}} \rangle
\delta \hat{z}^{\hat{J}}\biggr),
\label{gj2}
\end{eqnarray}
respectively.
While the expansion of $\hat{D}^A$ gives
\begin{eqnarray}
\delta \hat{D}^A &=&
\langle \hat{K}_\lambda \rangle (T^A)_\kappa^\lambda
\delta \hat{z}^\kappa
\nonumber \\
&~&+ (\langle \hat{K}_{{\hat{I}}\lambda} \rangle \delta
\hat{z}^{\hat{I}} + \langle \hat{K}_{\hat{I}} \rangle \delta
\hat{U}_{\lambda}^{\hat{I}})
(T^A)_\kappa^\lambda \langle z^\kappa \rangle,
\label{D1}\\
\delta^2 \hat{D}^A &=&
\langle \hat{K}_\lambda \rangle (T^A)_\kappa^\lambda
(\delta^2 \hat{z}^\kappa + \delta(\hat{U}^{-1})_{\hat{J}}^\kappa
\delta \hat{z}^{\hat{J}})
\nonumber \\
&~&+ ( \langle \hat{K}_{{\hat{I}}\lambda} \rangle \delta^2
\hat{z}^{\hat{I}} + \langle \hat{K}_{{\hat{I}}{\hat{J}}\lambda}
\rangle \delta \hat{z}^{\hat{I}} \delta \hat{z}^{\hat{J}}
\nonumber \\
&~&~~~~ + \langle \hat{K}_{\hat{I}} \rangle \delta^2
\hat{U}_{\lambda}^{\hat{I}}) (T^A)_\kappa^\lambda
\langle z^\kappa \rangle
\nonumber \\
&~&+ ( \langle \hat{K}_{{\hat{I}}\lambda} \rangle
\delta \hat{z}^{\hat{I}}
+\langle \hat{K}_{\hat{I}} \rangle \delta
\hat{U}_{\lambda}^{\hat{I}})
(T^A)_\kappa^\lambda \delta \hat{z}^\kappa .
\label{D2}
\end{eqnarray}
The expansions of the stationary conditions
$\partial V/\partial z^K=0$ and
$\partial V/\partial z^A=0$ give
\begin{eqnarray}
\langle \hat{W} \rangle_{KL} \langle
(\hat{K}^{-1})^{L{\bar{\mu}}} \rangle
\delta \widehat{{\cal G}}_{\bar{\mu}} &=& 0,
\label{stK1}\\
\langle \hat{W} \rangle_{KL} \langle
(\hat{K}^{-1})^{L{\bar{\mu}}} \rangle
\delta^2 \widehat{{\cal G}}_{\bar{\mu}}
&=&
- \delta \widehat{{\cal G}}_{\bar{\mu}} \langle
(\hat{K}^{-1})^{{\bar{\mu}}\lambda} \rangle
\langle \hat{W}_{\lambda\sigma K} \rangle \delta \hat{z}^{\sigma}
+ const.
\label{stK2}
\end{eqnarray}
and
\begin{eqnarray}
\langle Re f_{\alpha\beta}^{-1} \rangle \langle
(\hat{z} T^\alpha)^{\bar{\mu}} \rangle \langle
\hat{K}_{A\bar{\mu}} \rangle \delta \hat{D}^\beta &=& 0,
\label{stA1}\\
\langle Re f_{\alpha\beta}^{-1} \rangle \langle
(\hat{z} T^\alpha)^{\bar{\mu}} \rangle \langle
\hat{K}_{A\bar{\mu}} \rangle \delta^2 D^\beta &=&
E \delta \widehat{{\cal G}}_{\bar{\mu}}
\langle (K^{-1})^{{\bar{\mu}}\lambda} \rangle
\langle \hat{W}_{\lambda \sigma A} \rangle
\delta \hat{z}^\sigma \\
\nonumber
&~&+ const.,
\label{stA2}
\end{eqnarray}
respectively.
Here $E \equiv \langle exp(K/M^2) \rangle$.
{}From Eqs.~(\ref{gK1}), (\ref{gA1}),
(\ref{gj1}) and (\ref{stK1}), we find $\delta \hat{z}^K =0$ by
using $\langle \delta \hat{z}^K \rangle =0$.
Eq.~(\ref{stK2}) gives the solution for $\delta^2
\widehat{{\cal G}}_{\bar{K}}$ as
\begin{eqnarray}
\delta^2 \widehat{{\cal G}}_{\bar{K}} &=&
\langle \widehat{{\cal G}}_{\bar{K}} \rangle
-\langle (\hat{K}_{{\bar{K}}L}) \rangle
\langle \hat{W}^{-1} \rangle^{KL}
\delta \widehat{{\cal G}}_{\bar{\mu}} \langle
(K^{-1})^{{\bar{\mu}}M} \rangle \langle \hat{W}_{MKl}
\rangle \delta \hat{z}^l
\label{<gK>}
\end{eqnarray}
where a constant factor of $\delta^2 \widehat{{\cal G}}_{\bar{K}}$
is denoted as $\langle \widehat{{\cal G}}_{\bar{K}} \rangle$. From
Eqs.~(\ref{off2}), (\ref{D1}) and (\ref{stA1}),
we find $\delta \hat{D}^A = 0$ and $\delta \hat{z}^A = 0$.
By using the relations $\langle \hat{W}_{ABk} \rangle
=O(m_{3/2}/M_X)$ and
$\langle \hat{W}_{ABi} \rangle=O(m_{3/2}/M)$,
we can show that $\delta^2 \hat{D}^A$ is a
constant independent of the light fields.
Therefore we will denote it by $\langle \hat{D}^A \rangle$.
Now it is straightforward to calculate the scalar potential
${\cal V}^{eff}$ in the low-energy effective theory by
substituting the solutions of the stationary conditions
for the heavy fields.
The result can be compactly expressed if we define the effective
superpotential $\widehat{\cal W}_{eff}$ as
\begin{eqnarray}
\widehat{\cal W}_{eff} (z) &=&
{1 \over 2!}\hat{\mu}_{kl} \delta \hat{z}^k \delta \hat{z}^l
+ {1 \over 3!}\hat{h}_{klm} \delta \hat{z}^k \delta \hat{z}^l
\delta \hat{z}^m ,
\label{calWeff}
\end{eqnarray}
where
\begin{eqnarray}
\hat{\mu}_{kl} &\equiv& E^{1/2}\biggl(\langle
\hat{W}_{kl} \rangle + {\langle \hat{W} \rangle \over M^2}
\langle \hat{K}_{kl} \rangle
- \langle \hat{K}_{kl\bar{i}} \rangle \langle
(\hat{K}^{-1})^{{\bar{i}}j} \rangle
\delta \hat{\cal G}_j \biggr)
\nonumber \\
&~&~~~~ + (m^{'''}_{3/2})_{kl} ,
\label{hat-mu}\\
\hat{h}_{klm} &\equiv& E^{1/2} \langle \hat{W}_{klm} \rangle.
\label{hat-h}
\end{eqnarray}
Then we can write down the scalar potential of effective
theory as\footnote{
Here we omitted the terms irrelevant to the gauge non-singlet
fields $\delta \hat{z}^{\hat{k}}$ and the terms whose magnitudes
are less than $O(m_{3/2}^4)$.}
\begin{eqnarray}
{\cal V}^{eff}&=& {\cal V}_{SUSY}^{eff} + {\cal V}_{Soft}^{eff},
\label{calVeff}
\\
{\cal V}_{SUSY}^{eff} &=&
\frac{\partial \widehat{\cal W}^*_{eff}}{\partial \hat{z}^{\bar{k}}}
\langle (\hat{K}^{-1})^{{\bar{k}}l} \rangle
\frac{\partial \widehat{\cal W}_{eff}}{\partial \hat{z}^l}
+ {1 \over 2}g_a^2 (\langle \hat{K}_{k\bar{l}} \rangle
\hat{z}^{\bar{l}} (T^a)^k_l \hat{z}^l)^2,
\label{calVeffSUSY}
\\
{\cal V}_{Soft}^{eff}&=& A \widehat{\cal W}_{eff}
+ B_{\bar{k}}(\hat{z})_{eff}
\langle (\hat{K}^{-1})^{{\bar{k}}l} \rangle
\frac{\partial \widehat{\cal W}_{eff}}{\partial \hat{z}^l}
+ {\it H.c.}
\nonumber \\
&~&+ B_{\bar{k}}(\hat{z})_{eff}
\langle (\hat{K}^{-1})^{{\bar{k}}l} \rangle
{B_l(\hat{z})}_{eff} + C(\hat{z})_{eff}
\nonumber \\
&~& + \Delta \widehat{\cal V} + \Delta {\cal V}'^{(F)} ,
\label{calVeffsoft}
\end{eqnarray}
where $\Delta \widehat{\cal V} + \Delta {\cal V}'^{(F)}$
is a sum of contributions such as
\begin{eqnarray}
\Delta \widehat{\cal V} &=& \Delta \widehat{\cal V}_0^{(F)}
+ \Delta \widehat{\cal V}_0^{(D)} + \Delta
\widehat{\cal V}_1^{(F)} + \Delta \widehat{\cal V}_1^{(D)} ,
\label{DeltacalV}
\\
\Delta \widehat{\cal V}_0^{(F)} &\equiv&
E\{- \delta^2 \widehat{{\cal G}}_{\bar{K}} \langle
(\hat{K}^{-1})^{{\bar{K}}L} \rangle
\delta^2 \widehat{{\cal G}}_L
+\delta^2 \widehat{{\cal G}}_{\bar{A}} \langle
(\hat{K}^{-1})^{{\bar{A}}B} \rangle
\delta^2 \widehat{{\cal G}}_B
\nonumber \\
&~&~~~ +\delta \widehat{{\cal G}}_{\bar{A}} \langle
(\hat{K}^{-1})^{{\bar{A}}B} \rangle
\delta^{3'} \widehat{{\cal G}}_B + H.c.
\nonumber \\
&~&~~~ + \delta^2 \widehat{{\cal G}}_{\bar{\kappa}}
\langle (\hat{K}^{-1})^{{\bar{\kappa}}L}
\rangle \biggl(\langle \hat{W}_{Lk} \rangle \delta \hat{z}^k +
{1 \over 2}\langle \hat{W}_{Lkl} \rangle \delta \hat{z}^{k}
\delta \hat{z}^{l} + \cdots \biggr)
\nonumber \\
&~&~~~ + H.c.\} ,
\label{DeltacalV(F)0}
\\
\Delta \widehat{\cal V}_0^{(D)} &\equiv&
\langle Ref_{AB}^{-1} \rangle \langle \hat{D}^A \rangle
\langle \hat{K}_{{\hat{I}}\lambda} \rangle \delta
\hat{z}^{\hat{I}}
(T^B)_\kappa^\lambda \delta \hat{z}^\kappa ,
\label{DeltacalV(D)0}
\\
\Delta \widehat{\cal V}_1^{(F)} &\equiv&
(m^{*'''}_{3/2})_{\bar{K}\hat{k}} \delta \hat{z}^{\hat{k}}
\langle (\hat{K}^{-1})^{{\bar{K}}l} \rangle
\biggl(\frac{\partial \widehat{\cal W}_{eff}}
{\partial \hat{z}^l}-(m^{'''}_{3/2})_{lm}\delta \hat{z}^{m}
\nonumber \\
&~&~~~~ +(m_{3/2}+m_{3/2}^{''})_{l\bar{m}}\delta \hat{z}^{\bar{m}}
\biggr) +{\it H.c.}
\nonumber \\
&~& + E\{\delta \widehat{{\cal G}}_{\bar{K}} \langle
(\hat{K}^{-1})^{{\bar{K}}B} \rangle
\delta^{3'} \widehat{{\cal G}}_B + H.c.
+\delta \widehat{{\cal G}}_{\bar{\kappa}}
\delta^{2'} (\hat{K}^{-1})^{{\bar{\kappa}}\lambda}
\delta \widehat{{\cal G}}_{\lambda}
\nonumber \\
&~& +\delta \widehat{{\cal G}}_{\bar{\kappa}}
\delta (\hat{K}^{-1})^{{\bar{\kappa}}K}
\delta^{2'} \widehat{{\cal G}}_K + H.c.\} ,
\label{DeltacalV(F)1}
\\
\Delta \widehat{\cal V}_1^{(D)} &\equiv&
\langle Ref_{AB}^{-1} \rangle \langle \hat{D}^A \rangle
\langle \hat{K}_{{\hat{I}}{\hat{J}}\lambda} \rangle \delta
\hat{z}^{\hat{I}} \delta \hat{z}^{\hat{J}} (T^B)_\kappa^\lambda
\langle {z}^\kappa \rangle ,
\label{DeltacalV(D)1}
\\
\Delta {\cal V}'^{(F)} &=& E[Const.]^L
{1 \over 2} \langle \hat{W}_{Lkl} \rangle
\delta \hat{z}^{k} \delta
\hat{z}^{l} + H.c. ,
\label{DeltacalV'(F)}
\\
E[Const.]^L &\equiv& E [ \delta \hat{\cal G}_{\hat{I}}
\delta(\hat{U})^{\hat{I}}_{\bar{I}}
\langle (\hat{K}^{-1})^{{\bar{I}}L} \rangle
+ \delta \hat{\cal G}_{\bar{I}} \langle
(\hat{K}^{-1})^{{\bar{I}}J} \rangle \delta(\hat{U})^{L}_J
\nonumber \\
&~& + \delta \hat{\cal G}_{\bar{I}}(\langle
(\hat{K}^{-1})^{{\hat{J}}L} \rangle
\delta(\hat{U}^{-1})_{\hat{J}}^{\bar{I}}
+ \langle (\hat{K}^{-1})^{{\hat{I}}{\bar{I}}} \rangle
\delta(\hat{U}^{-1})^L_{\hat{I}}) ].
\label{constL}
\end{eqnarray}
The quantities with a prime such as $\delta^{3'}
\widehat{{\cal G}}_B$ mean that the terms proportional to
$\delta^{2} \hat{z}^{\hat{I}}$ are omitted.
The ellipses in Eq.~(\ref{DeltacalV(F)0}) represent other terms
in $\delta^2 \widehat{\cal G}_K - \langle \hat{W}_{KL} \rangle
\delta^2 \hat{z}^L$. (Refer Eq.~(\ref{gK2}).)
The soft SUSY breaking parameters $A$, $B_{\bar{k}}(z)_{eff}$
and $C(\hat{z})_{eff}$ are given as
\begin{eqnarray}
A &=& m_{3/2}^{\ast '} - 3m_{3/2}^{\ast},
\label{Aagain}\\
B_{\bar{k}}(\hat{z})_{eff} &=& (m_{3/2}^{\ast}
+ m_{3/2}^{\ast ''}+m_{3/2}^{\ast '''})_{{\bar{k}}l} \delta
\hat{z}^l,
\label{Bkeff}\\
C(\hat{z})_{eff} &=&
E \delta \hat{\cal G}_{\bar{i}} \langle
(\hat{K}^{-1})^{\bar{i}j} \rangle
\biggl({1 \over 3!}\langle \hat{W}_{jIJJ'} \rangle \delta \hat{z}^I
\delta \hat{z}^J \delta \hat{z}^{J'}
+ {\langle \hat{W} \rangle \over M^2}
\delta^{2'}\hat{K}_j \biggr) + H.c.
\nonumber \\
&~&+ E\biggl(\delta \hat{\cal G}_{\bar{i}}
\delta^{2'}(\hat{K}^{-1})^{\bar{i}j}
\delta \hat{\cal G}_{j} + {\langle V \rangle \over M^2}
\delta^{2'} \hat{K} \biggr)
\nonumber \\
&~& - (m_{3/2}^{*'''})_{l\bar{l}} \langle
(\hat{K}^{-1})^{k\bar{l}} \rangle
(m_{3/2}^{'''})_{k{\bar{k}}} \delta \hat{z}^{\bar{k}} \delta
\hat{z}^l
\nonumber \\
&~& - (m_{3/2}^{'''})_{kl} \langle (\hat{K}^{-1})^{k\bar{l}} \rangle
(m_{3/2}^{*'''})_{\bar{k}\bar{l}} \delta \hat{z}^{\bar{k}} \delta
\hat{z}^l
\nonumber \\
&~& - \{ (m_{3/2}^{'''})_{ml}
\langle (\hat{K}^{-1})^{m\bar{k}} \rangle
(m_{3/2}^*+ m_{3/2}^{*''})_{{\bar{k}}k}
\delta \hat{z}^{k} \delta \hat{z}^{l} + H.c. \}
\nonumber \\
&~& + A \biggl[ E^{1/2}\biggl( {\langle \hat{W} \rangle \over M^2}
\langle \hat{K}_{kl} \rangle
- \langle \hat{K}_{kl\bar{i}} \rangle \langle
(\hat{K}^{-1})^{{\bar{i}}j} \rangle
\delta \hat{\cal G}_j \biggr)
\nonumber \\
&~&~~~ + (m_{3/2}^{'''})_{kl} \biggr] \delta \hat{z}^k \delta
\hat{z}^l ,
\label{Ceff}
\end{eqnarray}
where
\begin{eqnarray}
(m_{3/2})_{k\bar{l}} &=& E^{1/2}{\langle \hat{W} \rangle \over M^2}
\langle \hat{K}_{k{\bar{l}}} \rangle,
\label{mkl*}\\
m_{3/2}^{'} &=& E^{1/2}{\langle \hat{K}_{\bar{i}} \rangle \over M^2}
\langle (\hat{K}^{-1})^{{\bar{i}}j} \rangle \delta\hat{\cal G}_j,
\label{m'}\\
(m_{3/2}^{''})_{k\bar{l}} &=& -E^{1/2}
\langle \hat{K}_{k\bar{l}\bar{i}} \rangle \langle
(\hat{K}^{-1})^{{\bar{i}}j} \rangle \delta\hat{\cal G}_j,
\label{m''}\\
(m_{3/2}^{'''})_{\kappa\bar{l}} &=& -E^{1/2}
\langle \hat{K}_{\kappa \bar{l}\bar{A}} \rangle \langle
(\hat{K}^{-1})^{{\bar{A}} \lambda} \rangle \delta
\widehat{{\cal G}}_{\lambda},
\label{m'''kl*}\\
(m_{3/2}^{'''})_{\kappa l} &=& -E^{1/2}
\langle \hat{K}_{\kappa l \bar{A}} \rangle \langle
(\hat{K}^{-1})^{{\bar{A}}\lambda} \rangle \delta
\widehat{{\cal G}}_{\lambda}.
\label{m'''kl}
\end{eqnarray}
The $\Delta \hat{\cal V}^{(D)}_0$ and $\Delta \hat{\cal V}^{(D)}_1$
come from the $D$-term of the heavy gauge sector and are referred
to as the $D$-term contributions,
while the others are called the $F$-term contributions.
We should consider the renormalization effects for the soft SUSY
breaking parameters and diagonalize the scalar mass matrix
$V_{\hat{k}\hat{l}}$ to derive the weak scale SUSY spectrum,
which is expected to be measured in the near future.
\section{Features of the Effective Lagrangian}
The effective theory obtained in the previous section
has some excellent features.
We discuss two topics.
\subsection{Chirality Conserving Mass}
\label{subsec:mass-terms}
We discuss a {\em chirality-conserving} mass term
$(m^2)_{k\bar{l}}$, namely the
coefficient of $\delta \hat{z}^k \delta \hat{z}^{\bar{l}}$.
They are easily extracted from ${\cal V}_{Soft}^{eff}$ and
given by
\begin{eqnarray}
(m^2)_{k\bar{l}} &=& (m^2_0)_{k\bar{l}}
+ (\Delta \widehat{\cal V}_0)_{k\bar{l}}
+ (\Delta \widehat{\cal V}_1)_{k\bar{l}} ,
\label{m2_kl*}
\\
(m^2_0)_{k\bar{l}} &\equiv&
{\partial \over \partial \hat{z}^k}B_{\bar{m}}(\hat{z})_{eff}
\langle (\hat{K}^{-1})^{m\bar{m}} \rangle
{\partial \over \partial \hat{z}^{\bar{l}}}B_{m}(\hat{z})_{eff}
\nonumber \\
&~&+ {\partial^2 \over \partial \hat{z}^k \partial
\hat{z}^{\bar{l}}} C(\hat{z})_{eff} ,
\label{m20_kl*}
\\
(\Delta \widehat{\cal V}_0)_{k\bar{l}} &\equiv&
{\partial^2 \over \partial \hat{z}^k \partial \hat{z}^{\bar{l}}}
\Delta \widehat{\cal V}_0^{(F)}
+ \langle Ref^{-1}_{AB} \rangle \langle \hat{D}^A \rangle
\langle \hat{K}_{{\bar{l}}\lambda} \rangle (T^B)^{\lambda}_k ,
\label{DeltaV0_kl*}
\\
(\Delta \widehat{\cal V}_1)_{k\bar{l}} &\equiv&
{\partial^2 \over \partial \hat{z}^k \partial \hat{z}^{\bar{l}}}
\Delta \widehat{\cal V}_1^{(F)}
+ 2 \langle Ref^{-1}_{AB} \rangle \langle \hat{D}^A \rangle
\langle \hat{K}_{k{\bar{l}}\lambda} \rangle
(T^B)^{\lambda}_{\kappa} \langle z^{\kappa} \rangle .
\label{DeltaV1_kl*}
\end{eqnarray}
The term $(m^2_0)_{k\bar{l}}$ is present before the
heavy sector is integrated out and so
it respects the original unified gauge symmetry.
On the other hand, other terms coming from
$\Delta \widehat{\cal V}$ can pick up effects of
the symmetry breaking.
The last terms in Eqs.~(\ref{DeltaV0_kl*}) and (\ref{DeltaV1_kl*})
are the $D$-term contributions.
We discuss the conditions of those existence.
The non-zero VEV of the $D$-term is allowed
for a $U(1)$ factor, {\it i.e.} a diagonal generator from
the gauge invariance.
And the $D$-term for an unbroken generator cannot have its VEV.
Thus it can arise when the rank of the
gauge group is reduced by the gauge symmetry breaking.
The $D$-term contribution is proportional to the charge of
the broken $U(1)$ factor and gives mass splittings within
the same multiplet in the full theory.
We can rewrite $\delta^2 \hat{D}^A= \langle \hat{D}^A \rangle$ as
\begin{eqnarray}
\langle \hat{D}^A \rangle =
2(M_{V}^{-2})^{AB} E
\delta \widehat{\cal G}_{\kappa} \delta \widehat{\cal G}
_{\bar{\lambda}}
\{ G_{\bar{\mu}}^{\kappa\bar{\lambda}} (\hat{z} T^B)^{\bar{\mu}}
+ G^{\bar{\mu}\kappa} (T^B)_{\bar{\mu}}^{\bar{\lambda}} \}
\label{<hatD>}
\end{eqnarray}
by using the gauge invariance.
We can see that the VEVs vanish up to $O(m_{3/2}^4/M_X^2)$
when the K\"ahler potential has the minimal structure.
Hence the sizable $D$-term contribution can appear only when
the K\"ahler potential has a non-minimal structure.
The other terms in Eqs.~(\ref{DeltaV0_kl*}) and
(\ref{DeltaV1_kl*}) are related to the $F$-terms.
They can be neglected in the case that the superpotential
couplings are weak and the $R$-parity conservation is assumed.
Therefore phenomenologically the $D$-term contribution
to the scalar masses is important to probe SUSY-GUT models
because it can give an additional contribution to squarks,
sleptons and Higgs bosons\cite{Sfermion}.
\subsection{Specific Case}
Finally we discuss the relation between
our result and that in Ref.\cite{KMY2}.
For later convenience,
we list up features in the approach of Ref.\cite{KMY2}.
\begin{enumerate}
\item The starting theory is a unified theory obtained by
taking the flat limit of SUGRA with a certain type of total
K\"ahler potential,
so the terms of order $m_{3/2}^4(M_X/M)^n$ are neglected.
Since the unification scale $M_X$ is now believed to be lower
than the gravitational scale $M$ from LEP data\cite{LEP},
this procedure can be justified in such a model.
However it will be important when higher order corrections
are to be considered.
Then we must incorporate threshold effects and loop effects.
\item The scalar fields have canonical kinetic terms.
It was assumed that the SUSY fermion mass matrix and the
kinetic function can be diagonalized simultaneously, i.e.,
the relation $\langle \hat{K}_{\kappa{\bar{\lambda}}} \rangle
= \delta_{\kappa{\bar{\lambda}}}$ is imposed.
\item The {\it Hidden} assumption on superpotential was
taken because it was purposed to discuss consequences
independent of the details of each models.
The stability of gauge hierarchy is automatically guaranteed
under this assumption.
\item The heavy-light mixing, in general, can occur after soft
SUSY breaking terms are incorporated.
Then we must re-define the scalar fields by diagonalizing
the mass matrix.
It was assumed that there is no heavy-light mixing
after SUSY breaking.
\end{enumerate}
We shall derive the previous one ${\cal V}^{eff(non)}$ from
our scalar potential ${\cal V}^{eff}$
by refering to the list.
\begin{enumerate}
\item When we take the limit $M_X/M \longrightarrow 0$, we
find that some terms vanish.
For example, $(m_{3/2}^{'''})_{\kappa\bar{l}}$ and
$(m_{3/2}^{'''})_{\kappa l}$ vanish
and
\begin{eqnarray}
\Delta \widehat{\cal V}_1 \longrightarrow
E\{\delta \hat{{\cal G}}_{\bar{K}} \langle
(\hat{K}^{-1})^{{\bar{K}}B} \rangle \delta^{3'} \hat{{\cal G}}_B
+ H.c.\}.
\end{eqnarray}
\item Next we impose the condition
$\langle \hat{K}_{\kappa{\bar{\lambda}}} \rangle
= \delta_{\kappa{\bar{\lambda}}}$.
Then $\Delta \widehat{\cal V}_1$ and some other terms vanish.
\item Further we take the {\it hidden} ansatz
$\langle \hat{W}_{j...k...} \rangle = 0$.
Then the trilinear coupling constant is reduced to
\begin{eqnarray}
A E^{1/2} \langle \hat{W}_{klm} \rangle + {1 \over 2}
(m_{3/2}^* + m_{3/2}^{*''})_{k\bar{k}}
\langle (\hat{K}^{-1})^{{\bar{k}}n} \rangle \langle
\hat{W}_{nlm} \rangle.
\end{eqnarray}
\item When we take a model with no heavy-light mixing,
$\Delta {\cal V}'^{(F)}$ does not exist.
We can find an ansatz for the K\"ahler potential
that the heavy-light mixing does not occur
in the gauge non-singlet sector after taking the flat limit.
For example, the ansatz
\begin{eqnarray}
K &=& K^{({\cal H})}(z^{{\cal H}}, z_{{\cal H}}^*
;\tilde{z}, \tilde{z}^*) +
K^{({k})}(z^{k}, z_{k}^*
;\tilde{z}, \tilde{z}^*)
\end{eqnarray}
fulfills our requirement.
\end{enumerate}
We find that ${\cal V}^{eff}$ reduces to $V^{eff(non)}$
after the above procedures.
\section{Conclusions}
We have derived the low-energy effective Lagrangian from
SUGRA with non-minimal structure and unified gauge symmetry
in model-independent manner.
The starting SUGRA is more general one than those considered
before.
The total K\"alher potential has a non-minimal structure
based on the hidden sector SUSY breaking scenario.
We have distinguished between the scales $M_X$ and $M$.
It is important to investigate its consequences at low-energy
because the non-minimal SUGRA appears naturally in many
circumstanses.
For example, SSTs lead to the non-minimal SUGRA effectively.
Even if SUGRA have the minimal structure at the tree level, it
can get renormalized and as a result, in general,
become non-minimal.
We have calculated the scalar potential by
taking the flat limit and integrating out the heavy sector.
The result is summarized in Eqs.~(\ref{calWeff})
--(\ref{m'''kl}).
We found new contributions to the soft terms
reflected to the non-minimality and the breaking of
unified gauge symmetry.
In particular, the sizable $D$-term contributions
generally exist in the scalar masses when the
rank of the gauge group is reduced by the gauge symmetry
breaking and the K\"alher potential has a non-minimal structure.
Its phenomenological implications were discussed
in Ref.\cite{Sfermion}.
Another important point is the gauge hierarchy problem.
Many SUSY-GUT models achieve the small
Higgs doublet masses by a fine-tuning of the parameters in the
superpotential.
If the SUSY breaking due to the hidden field condensations is
turned on, a SUSY breaking Higgs mass term can become heavy
and the weak scale can be destabilized.
We have shown that the masses of light fields remain at the
weak scale if the couplings of hidden-sector fields to
visible-sector fields in the superpotential
satisfy certain requirements.
We have derived the results in Ref.\cite{KMY2} by taking
some limit and conditions.
We also have studied the SUGRA with Fayet-Iliopoulos $D$-term
and derived the low-energy effective theory.
It is believed that the measurements of SUSY spectrum at
the weak scale can be useful in probing physics at SUSY-GUT
and/or SUGRA, if the SUSY breaking scenario through the
gauge-singlet sector in SUGRA is realized in nature.
Hence the precision measurements should be carried out by the
colliders in the near future.
\section*{Acknowledgements}
The author is grateful to H.~Murayama, H.~Nakano and I.~Joichi
and especially M.~Yamaguchi for useful discussions.
This work is supported by the Grant-in-Aid for
Scientific Research ($\sharp$07740212) from
the Japanese Ministry of Education, Science and Culture.
|
\section{Introduction}
One of the mysteries of heavy ion physics at Brookhaven National
Laboratory's AGS is: {\it If hadronic cascade event simulators like
RQMD \cite{rqmd} and ARC \cite{arc} produce energy densities
approaching 2 GeV/fm$^3$, yet agree with experiment, where is
the quark--gluon plasma?} After all, numerous estimates of the
onset of quark--gluon plasma agree that it should occur at about
that energy density, and if there is a first order phase transition,
then the onset of the mixed phase would occur at an even lower density.
One possibility is that no phase transition occurs even at these
high densities, but it is difficult to understand how composite
objects like hadrons can overlap so strongly in position space
without the matter undergoing some qualitative change in character.
A second possibility is that the distribution of observed hadrons
in the final state is insensitive to the dynamics of the matter
when it is most hot and dense. (Unfortunately there are no
measurements of direct photons or dileptons at the AGS which might
probe this stage of the collision.) There is some evidence for
this which comes from artificially modifying hadronic cross sections
at high density \cite{pang}. It may be understood by recognizing
that once a system reaches local thermal equilibrium it is basically
irrelevant how it got there.
Recently we proposed a third possibility \cite{us}: {\it Most collisions
at AGS energies produce superheated hadronic matter and are describable
with hadronic cascade simulators, but in rare events a droplet of
quark--gluon plasma is nucleated which converts most of the matter to
plasma.} We estimated the probability of
this to occur, using homogeneous nucleation theory, to be on the
order of once every 100 to 1000 central collisions of large nuclei.
Our estimate was based on the probability that thermal
fluctuations in a homogeneous superheated hadronic gas would produce
a plasma droplet, and that this droplet was large enough to overcome
its surface free energy to grow. In this paper we consider another
source of plasma droplet production which is essentially one of
nonthermal origin. Specifically, we estimate the probability
that a collision occurs between two highly energetic incoming nucleons,
one from the projectile and one from the target,
that this collision would have produced many pions if it had occurred
in vacuum, but because it occurs in the hot and dense medium
its collision products are quark and gluon fields which make a small
droplet of plasma. Although there is a large uncertainty in our
estimates, we find that this inhomogeneous nucleation of plasma may be
more probable than homogeneous nucleation by one to two orders of magnitude.
In this paper we also consider the problem of observation of the
effects of nucleation of plasma in rare events. We are guided by
observations of multiplicity distributions in $p\bar{p}$ collisions
at the CERN and Fermilab colliders. In those distribution, one
sees a shoulder developing at high multiplicity at an energy of
540 GeV, which turns into a noticeable bump at higher energies.
The real cause of this structure is not known, but may be due to
minijet production. If plasma is nucleated in some fraction of
central nucleus--nucleus collisions at the AGS, a similar structure
may develop.
\section{Kinetic Model of Hard Nucleon-Nucleon Collisions}
In this section we develop a simple kinetic model which allows us
to estimate the number of high energy nucleon-nucleon scatterings
occurring in the high density medium formed during a collision
between heavy nuclei. These scatterings occur when a projectile
nucleon penetrates the hot and dense matter to collide with
a target nucleon which has also penetrated the hot and dense matter.
The energy loss of the colliding nucleons must be taken into
account to obtain a reasonable estimate of the energy available
for meson production in the nucleon-nucleon collision.
To first approximation we can visualize the initial stage of a
heavy ion collision at the AGS in the nucleus-nucleus center-of-
momentum frame as two colliding Lorentz contracted disks.
See Figure 1. At time $t = 0$ they touch; subsequently they
interpenetrate, forming hot and dense matter in the region of overlap.
During this stage, additional matter streams into the hot zone
even as this zone is expanding along the beam axis. The nucleons
streaming in undergo scatterings with the hot matter already present,
degrading their longitudinal momentum and producing baryonic isobars
and/or mesons. Finally, at time $t_0 = L/2v\gamma$, all the
cold nuclear matter has streamed into the region of overlap,
and expansion and cooling begins. Here, $L$ is the nuclear thickness,
$v$ is the velocity in the center-of-momentum frame, and $\gamma$
is the associated Lorentz contraction factor. This is a very
simplified picture of the early stage of the collision, but it
seems to semi-quantitatively represent the outcome of both
the ARC and RQMD simulations \cite{rqmd,arc,us}.
We are interested in the possibility that an incoming projectile
nucleon suffers little or no energy loss during its passage
to the longitudinal point $z$ inside the hot and dense zone
where it encounters a target nucleon which also has suffered
little or no energy loss. The energy available in the ensuing
nucleon-nucleon collision, $\sqrt{s}$, can go into meson production.
Suppose that a large number of pions would be produced if the collision had
happened in free space. Clearly, the outgoing quark and gluon fields
cannot be represented as asymptotic pion and nucleon states immediately.
The fields must expand and become dilute enough to be called real
hadrons. If this collision occurs in a high energy density medium,
the outgoing quark-gluon fields will encounter other hadrons before
they can hadronize. It is reasonable to suppose that this ``star
burst" will actually be a seed for quark-gluon plasma formation
if the surrounding matter is superheated hadronic matter. We need
a semi-quantitative model of this physics.
A fundamental result from kinetic theory is that the number of
scattering processes of the type 1 + 2 $\rightarrow X$ is given by
\begin{equation}
N_{1+2 \rightarrow X}
= \int dt \int d^3 x \int \frac{d^3 p_1}{(2 \pi )^3} \, f_1 ({\bf x},
{\bf p}_1,t) \int \frac{d^3 p_2}{(2 \pi )^3} \, f_2 ({\bf x},{\bf p}_2,t)
\, v_{12} \, \sigma_{1+2 \rightarrow X}(s_{12}) \, .
\end{equation}
Here $v_{12}$ is a relative velocity,
\begin{equation}
v_{12} = \frac{\sqrt{(p_1 \cdot p_2)^2 -m_N^4}}{E_1 \, E_2} \, ,
\end{equation}
where $p_i$ denotes the four-momentum of nucleon $i$ and $E_i =\sqrt{
{\bf p}_i^2 + m_N^2}$ its energy. The $f_i$ are phase space
densities normalized such that the total number of nucleons of type
$i$ is
\begin{equation}
N_i^{\rm tot} = \int \frac{d^3x d^3p}{(2\pi)^3} f_i({\bf x,p},t) \, .
\end{equation}
A differential distribution in the variable $Y$ is obtained by replacing
$\sigma$ with $d\sigma/dY$.
For our purpose it is reasonable to represent the colliding
nuclei as cylinders with radius $R$ and thickness $L$. All the
action is along the beam axis. We assume that the phase space
distributions are independent of transverse coordinates $x$ and
$y$ and of transverse momentum. Integrating over the cross sectional
area of the nuclei, and counting only those collisions that occur
within the hot zone, yields
\begin{equation}
N_{1+2 \rightarrow X} = \pi R^2 \int_0^{t_0} dt \int_{-vt}^{vt} dz
\int \frac{dp_{1z} \, dp_{2z}}{(2 \pi)^2}
f_1 (z,p_{1z},t) \, f_2 (z,p_{2z},t) \, v_{12}
\, \sigma_{1+2 \rightarrow X}(s_{12}) \, .
\end{equation}
Here there is a change in notation: $f_i(z,p_{iz},t)/2\pi$ is the
probability per unit volume to find a nucleon $i$ with longitudinal
momentum $p_{iz}$ at longitudinal position $z$ at time $t$.
The integration limits on $z$ ensure that the collisions under
consideration really occur in the hot zone; see Figure 1.
The integration limits on $t$ mean that we only count those
collisions which occur before the system begins its cooling stage.
The depth in the hot zone to which nucleon 1 has penetrated is
$d_1 = (vt+z)/2$, and the depth to which nucleon 2 has penetrated
is $d_2 = (vt-z)/2$. We neglect the decrease in velocity of the
nucleons as they travel through the hot zone. This is an acceptable
approximation because in the end we are interested only in those
nucleons which suffer a small energy loss in traversing the hot
matter.
We construct the phase space distribution as follows:
\begin{verse}
$H(x,N)$ = probability that the nucleon has momentum fraction
$x$ after making N collisions;\\
$S(N,d)$ = probability that the nucleon has made $N$ collisions
after penetrating to a depth $d$;\\
$\sum_{N=0}^{\infty} H(x,N) S(N,d)$ = probability that the nucleon
has momentum fraction $x$ after penetrating to a depth $d$.
\end{verse}
The distribution functions are normalized to unity.
\begin{eqnarray}
\int_{0}^{1} \frac{dx}{x} \, H(x,N) &=& 1 \\
\sum_{N = 0}^{\infty} \, S(N,d) &=& 1
\end{eqnarray}
The phase space density of nucleon $i$ is then taken to be
\begin{equation}
\frac{dp_{zi}}{2\,\pi} \, f_i (z,p_{iz},t) =
\gamma \, n_0 \, \frac{dx_i}{x_i} \sum_{N_i = 0}^{\infty}
\, H(x_i,N_i) \, S(N_i,d_i) \, ,
\end{equation}
where $n_0$ is the average baryon density in a nucleus, about 0.145
nucleons/fm$^3$. As a check, we can compute the number of nucleons
which have entered the hot zone as a function of time.
\begin{equation}
N_i^{\rm part}(t) = \int \frac{d^3x dp_{iz}}{2\pi} f(z,p_{1z},t) \, \Theta
(d_i)
= 2\pi R^2 \gamma n_0 v t
\end{equation}
The step function fixes the limits on the $z$ integration. The
number of participating nucleons grows linearly with time, and
at time $t_0$ we get $N_i^{\rm part}(t_0) = \pi R^2 L n_0$, which
is the total number of nucleons in the nucleus.
The number of elementary nucleon-nucleon collisions can now be
expressed as
\begin{eqnarray}
N_{1+2 \rightarrow X} &=& \pi R^2 \gamma^2 n_0^2 \int_0^{t_0} dt
\int_{-vt}^{vt} dz \int_0^1 \frac{dx_1}{x_1} \int_0^1 \frac{dx_2}{x_2}
\, v_{12} \, \sigma_{1+2 \rightarrow X}(s_{12}) \nonumber \\
&\,& \sum_{N_1 = 0}^{\infty} \sum_{N_2 = 0}^{\infty} H(x_1,N_1)\,S(N_1,d_1)\,
H(x_2,N_2)\,S(N_2,d_2) \, .
\end{eqnarray}
Since the nucleons' velocities are antiparallel the velocity factor is
\begin{equation}
v_{12} = \frac{x_1 p_0}{\sqrt{x_1^2 p_0^2 + m_N^2}}
+ \frac{x_2 p_0}{\sqrt{x_2^2 p_0^2 + m_N^2}} \, ,
\end{equation}
where $p_0$ is the beam momentum in the center-of-momentum frame.
The survival function $S(N,d)$ is characterized by the mean free
path $\lambda$ of nucleons in the hot and dense hadronic matter.
For a dilute gas the inverse of the mean free path is the sum of
products of the cross section of the nucleon with the density
of objects it can collide with.
\begin{equation}
\lambda^{-1} = \sum_i \, n_i \sigma_i
\end{equation}
Average particle densities, including baryons and mesons,
were computed in ref. \cite{us} for the hot and dense matter
under consideration. A plot of the density as a function of
beam energy is shown in Figure 2. Assuming an average hadron-nucleon
cross section of 25 mb, we find $\lambda$ = 0.4 fm at a laboratory
beam energy of 11.6 GeV/nucleon. This is very short, and just
emphasizes the physics we discussed in the introduction concerning
hadronic matter versus quark-gluon plasma.
We assume that the collisions suffered by the nucleons are independent
and can be characterized by a Poisson distribution.
\begin{equation}
S(N,d) = \frac{1}{N!} \left( \frac{d}{\lambda}\right)^N
\exp{\left(-\frac{d}{\lambda}\right)}
\end{equation}
Here $d/\lambda$ is the average number of scatterings in a
distance $d$.
The invariant distribution function $H(x,N)$ describes the momentum degradation
of a nucleon propagating through the hot zone. This distribution
function was introduced in the evolution model of Hwa \cite{hwa}.
In this model the nucleon propagates on a straight line trajectory and
interacts with target particles contained within a tube with area
given by the elementary nucleon--nucleon cross section $\sigma_{NN}$.
Csernai and Kapusta \cite{evol} solved the resulting evolution
equations and found that the invariant distribution function in this model
is given by
\begin{eqnarray}
H(x,N) = x \sum_{n=1}^N \left( \begin{array}{c} N\\n \end{array} \right)
w^n (1-w)^{N-n}\,\, \frac{(-\ln{x})^{n-1}}{(n-1)!} +(1-w)^N \, \delta (x-1)
\, .
\end{eqnarray}
The $\delta$-function represents elastic and soft inelastic contributions
to the evolution of the nucleon through the matter.
The probability $w$ is the ratio of inelastic to total nucleon--
nucleon cross section. It corresponds to the probability that the nucleon
scatters inelastically and therefore drops out of the evolution described
by $H$; it is approximately 0.8 in free space. Csernai and Kapusta
found that it reduces to about 0.5 for nucleons propagating through a
nucleus. This value allowed them to obtain a good representation of data with
beam energies in the range of 6-405 GeV. In our case the nucleon is
propagating through hot and dense hadronic matter. We keep $w$ as a free
parameter since we don't know how the value of $w$ changes due to the thermal
excitations and the increased density.
We are interested in the number of pion-producing nucleon-nucleon collisions
with a relatively high center-of-momentum energy squared $s$. Our
basic result from this section is
\begin{eqnarray}
\frac{dN^{\rm hard}_{\rm in}}{ds} &=& \pi R^2 \gamma^2 n_0^2 \,
\sigma_{\rm in}(s)
\int_0^{t_0} dt \int_{-vt}^{vt} dz
\int_0^1 \frac{dx_1}{x_1} \int_0^1 \frac{dx_2}{x_2} \,
v_{12} \, \delta(s-s_{12}) \nonumber \\
&\,& \sum_{N_1 = 0}^{\infty} \sum_{N_2 = 0}^{\infty} H(x_1,N_1)\,
S(N_1,d_1(z,t))\,H(x_2,N_2)\,S(N_2,d_2(z,t)) \, .
\label{dncoll}
\end{eqnarray}
Here $\sigma_{\rm in}$ is the inelastic nucleon-nucleon cross section,
and $\sqrt{s_{12}}$ is the total energy in the nucleon-nucleon collision
where the nucleons have momentum fractions $x_1$ and $x_2$.
\section{Meson Production Cross Sections}
A phase transition to quark--gluon
plasma will become thermodynamically favorable if the
energy density is large enough. The corresponding phase
boundary in the temperature/chemical potential plane was explored in
\cite{us}. Until now we have only selected nucleon-nucleon scatterings
in which the total available energy $\sqrt{s}$ is large.
In addition, we need to specify what fraction of this energy
goes into meson production.
In this section we estimate the pion number distribution function
$P_n (s)$, which is the probability of producing $n$ pions in a
nucleon--nucleon collision in free space. The pion number
distribution function is linked to the cross section $\sigma_n$
for producing $n$ pions by
\begin{equation}
P_n (s) = \sigma_n (s)/ \sigma_{\rm in}(s) \,.
\label{topo}
\end{equation}
Given $P_n (s)$ we can estimate the number of nucleon-nucleon collisions
that would lead to the production of $n$ pions as
\begin{equation}
N_n = \int_{s_{\rm min}}^{4E_0^2} ds\,P_n (s) \,
\frac{dN_{\rm in}^{\rm hard}}{ds}\,.
\label{number}
\end{equation}
The lower limit of integration is fixed by kinematics and
the upper limit is determined by the beam energy.
We shall approximate the pion number distribution function
$P_n(s)$ with a binomial \cite{comment} and
choose the parameters of this binomial such that we have some rough
agreement with experiment \cite{topo}.
\begin{equation}
P_n (s) = \left( \begin{array}{c} n_{\rm max}\\n \end{array} \right)
\xi^n
(1-\xi)^{n_{\rm max}-n}
\label{bino}
\end{equation}
The maximum number of pions produced in a nucleon-nucleon collision
is determined by kinematics.
\begin{equation}
n_{\rm max} (s)= {\rm Integer} \left(\frac{\sqrt{s}-2m_N)}{m_{\pi}}\right)
\label{nc}
\end{equation}
The parameter $\xi$ is related to the mean multiplicity by
\begin{equation}
\xi(s) = \frac{\langle n \rangle}{n_{\rm max}}
= \frac{3}{n_{\rm max}}\,\left(\frac{1}{4} \langle n_{pp}^- \rangle
+\frac{1}{2} \langle n_{pn}^- \rangle
+\frac{1}{4} \langle n_{nn}^- \rangle\right)\, .
\label{q}
\end{equation}
Here $\langle n \rangle$ is the average pion multiplicity averaged over
$pp$, $pn$ and $nn$ collisions while $\langle n_{pp}^- \rangle$,
$\langle n_{pn}^- \rangle$ and $\langle n_{nn}^- \rangle$ represent
the average negative pion multiplicity in those collisions.
All average multiplicities are functions of $s$, of course. The
factor of 3 is due to isospin averaging.
Experimental data were compiled and parametrized in \cite{multi} as
\begin{eqnarray}
\langle n_{pp}^- \rangle &=& -0.41 + 0.79 F(s) \nonumber \\
\langle n_{pn}^- \rangle &=& -0.14 + 0.81 F(s) \nonumber \\
\langle n_{nn}^- \rangle &=& +0.35 + 0.77 F(s) \, .
\label{paras}
\end{eqnarray}
The function $F$ was introduced by Fermi \cite{fermi},
\begin{equation}
F(s) = \frac{(\sqrt{s}-2m_N)^{3/4}}{s^{1/8}}\, ,
\label{ferf}
\end{equation}
with $s$ measured in GeV$^2$.
The parametrizations in (\ref{paras}) describe the data rather well
except in the threshold region. We approximate the inelastic
nucleon--nucleon cross section $\sigma_{\rm in}$ by
the inelastic proton--proton cross section. A convenient parametrization
is given in \cite{topo},
\begin{equation}
\sigma_{\rm in} = 30.9 - 28.9\,p_L^{-2.46} - 0.835\,\ln{p_L}
+0.192\,\ln^2{p_L}\, ,
\label{sin}
\end{equation}
where $p_L$ is the laboratory momentum in GeV/c and the cross section
is in mb. This parametrization is good for $p_L > 0.968$ GeV/c.
The pion production cross sections, as described above, are displayed in
Figure 3. They have the right shapes and the right orders of magnitude
compared to data \cite{topo}. However, direct comparison is not possible.
First of all, data generally does not exist for final states with
$\pi^+$, $\pi^-$, and $\pi^0$. Usually, exclusive experiments can
only measure charged mesons or neutral mesons, not both. Secondly,
we have not been so sophisticated as to include vector mesons,
the $\eta$ meson, and kaons. For our purpose such sophistication is
probably not necessary. We care only about the probability that
a nucleon-nucleon collision leads to a significant amount of energy
release in the sense of conversion of initial kinetic energy to
meson mass. We are essentially basing our results on the total
inelastic cross section, the average meson multiplicity, kinematics,
and entropy. Our analysis would be better if we had a handle
on the width of the multiplicity distribution, averaged over the initial
state isospin and summed over the final state isospin.
\section{Star Burst Probabilities}
In this section we put together the ingredients developed in the
last two and compute the number of star bursts which may become
nucleation sites or seeds for plasma formation and growth.
The nucleon-nucleon collisions may be referred to as primary-primary,
primary-secondary, and secondary-secondary, depending on whether
the nucleons have scattered from thermalized particles in the
hot zone (secondary) or not (primary). The easiest contribution to
obtain is the primary-primary. All integrations and summations can
be done analytically with the result
\begin{equation}
\frac{dN^{\rm prim-prim}_{\rm in}}{ds} =
4\pi R^2 \sigma_{\rm in}(s) \left(\frac{\lambda \gamma n_0}
{w}\right)^2 \, \left[1-\left(1+w\,\frac{vt_0}{\lambda}\right)\,
\exp{\left(-\frac{vt_0}{\lambda}\right)}\right] \,
\delta (s-4E_0^2) \, .
\end{equation}
The formulas for the primary-secondary and secondary-secondary
contributions can be simplified to some extent but in the end
some summations remain which must be done numerically.
The number of nucleon-nucleon collisions as a function of $s$
are plotted in Figure 4. Both $w$ = 0.5 and 0.8 are shown;
there is little difference. The laboratory beam energy is 11.6
GeV per nucleon and the nuclei are gold. The spike represents
the delta function from primary-primary collisions. The contribution
from primary-secondary collisions falls from about 11 to 7 GeV$^{-2}$
as $s$ goes from 9 to 25 GeV. The contribution from secondary-secondary
collisions is almost negligible.
The pion multiplicity distribution arising from these hard collisions
is shown in Figure 5. It drops by more than nine orders of magnitude
in going from 6 pion production to 18 pion production. Typically there
is only one hard nucleon-nucleon collision leading to the production
of seven pions in a central gold-gold collision at this energy.
We are interested in the possibility that one of these star bursts
nucleates quark-gluon plasma. The precise criterion for this to
happen is not known. However, we can make some reasonable estimates.
In \cite{us} we estimated that a critical size plasma droplet at
these temperatures and baryon densities would have a mass of about
4 GeV. Any local fluctuation more massive than this would grow
rapidly, converting the surrounding superheated hadronic matter to
quark-gluon plasma. A similar estimate, based on the MIT bag model,
a simpler hadronic equation of state (free pion gas),
and with zero baryon density,
was obtained much earlier \cite{old}. Another estimate is obtained
by the argument that at these relatively modest beam energies most
meson production occurs through the formation and decay of baryon
resonances: $\Delta$, $N^*$, etc. The most massive observed resonances
are in the range of 2 to 2.5 GeV. Putting two of these in close
physical proximity leads to a mass of 4 to 5 GeV. We now need an
estimate of the number of pions this critical mass corresponds to.
Let us assume that each particle, nucleon and meson, carries away
a kinetic energy equal to one half its rest mass. If a particle
would have too great a kinetic energy then it might escape from
the nucleon-nucleon collision volume long before its neighbors and
so would not be counted in the rest mass of the local fluctuation.
Taking 4 GeV, dividing by 1.5, and subtracting twice the nucleon mass
leaves about 6 pion rest masses. So our most optimistic estimate
is that one needs a nucleon-nucleon collision which would have led
to 6 pions if it had occurred in free space. One might be less
optimistic and require the production of 8 or 10 pions instead.
In Figure 6 we show the total number $N_>$ of nucleon-nucleon collisions
which would lead to the production of at least $n_{\rm crit}$ pions.
We may view $n_{\rm crit}$ as the minimum number necessary to form
a nucleation site or plasma seed. If $n_{\rm crit}$ = 6 is the
relevant number then there are on average 7 such nucleon-nucleon
collisions per central gold-gold collision. If 8 or 10 are the
relevant multiplicities then there is only one such critical star
burst every 1 or every 25 central gold-gold collisions, respectively.
These numbers vary somewhat with $w$; the numbers quoted are averages.
Conservatively, we may conclude that the probability of at least
one plasma seed appearing via this mechanism is in the range of
1 to 100\% per central gold-gold collision at the highest energy
attainable at the AGS. These probabilities are about one to two
orders of magnitude greater than those estimated in \cite{us}
on the basis of thermal homogeneous nucleation theory.
\section{Consequences for the Multiplicity Distribution}
The results of the last section confirm the possibility of producing
quark--gluon plasma droplets in rare events at AGS. Once formed the droplets
grow rapidly due to the significant superheating of the hadronic matter. This
process was explored in~\cite{us} where it was found that the radii of such
droplets can reach $3-5$ fm. Since the phase transition
is occurring so far out of equilibrium we would expect a significant
increase in the entropy of the final state. This could be seen in the
ratio of pions to baryons, for example, or in the ratio of deuterons
to protons \cite{me}. Along with the increased entropy should come
a slowing down of the radial expansion due to a softening in the
matter, that is, a reduction in pressure for the same energy
density. Together, these would imply a larger source size and
a longer lifetime as seen by hadron interferometry \cite{scott}.
In this section we study one of the experimental ramifications in detail.
Specifically, we look at the charged particle multiplicity distributions and
investigate under what conditions one might be able to detect the rare events
from the structure of this distribution.
In Figure 7 we plot the ratio of entropy to total baryon number
$S/B$ for the hadronic and quark--gluon plasma phase for fixed beam
energies. Fixed beam energy means that initially both the energy density
and the baryon number density of the system is given which then determine
the corresponding entropies via the equation of state. We use the equation
of state discussed in~\cite{us} for all further calculations.
It is helpful to consider two extreme and opposite scenarios. Either the
matter stays all the time in the hadronic phase, or the matter has
been completely converted to quark--gluon plasma by the time $t_0$ and
only hadronizes later.
The difference of the entropies produced in these two scenarios is given by the
difference of the two curves in Figure 7. It represents an upper limit on the
additional number of pions produced. Since the temperature is comparable to
or larger than the pion mass the excess entropy is proportional to the
maximum number of excess pions
\begin{eqnarray}
3 \frac{\Delta N_{\rm -}}{B} = \frac{1}{3.6} \frac{\Delta S}{B}\, .
\label{npi}
\end{eqnarray}
The number of additional negatively charged pions per baryon
$\Delta N_{\rm -}/B$ is linearly related to the entropy difference $\Delta S$
determined from Figure 7.
The result is shown in Figure 8 for central Au + Au collisions. At beam
energies of $11.6$ GeV/A we produce $0.33$ additional negatively charged pions
per participating baryon. This is an upper limit, and in reality we
would expect less.
These additional mesons might be visible in the charged
particle multiplicity distribution which would have the form
\begin{eqnarray}
P_n = (1-q) \, P_{\; n}^{\rm had} (N_{\rm had})
+ q \, P^{\rm qg}_{\; n} (N_{\rm qg})\, .
\label{double}
\end{eqnarray}
Here $q$ is the probability of finding a central event in which plasma
is formed, $P_{\; n}^{\rm had}$ is the multiplicity distribution
for purely hadronic events with mean $N_{\rm had}$, and $P^{\rm qg}_{\; n}$
is the multiplicity distribution for events in which a plasma
was formed with mean $N_{\rm qg}$.
Experimentally one would expect to see a bump in $P_n$ at larger values
of $n$. A structure like that was found in charged particle multiplicity
distributions in $p\bar{p}$ collisions at the CERN \cite{ua5,fuglesang} and
Fermilab \cite{cdf,e735} colliders. For energies larger then 540 GeV a
shoulder develops in the multiplicity distribution, becoming more pronounced
as the beam energy increases. It is assumed that this structure is due
to the onset of minijets. It is definitely an indication of new physics.
In Figure 9 we plot the charged particle multiplicity distribution for
$p\bar{p}$ collisions at $\sqrt{s} = 900$ GeV from the UA5
collaboration~\cite{ua5}.
For energies less then 500 GeV it was found that the distribution could be well
described by a negative binomial distribution of the form
\begin{eqnarray}
P_n (\bar{n}, k) =
\left( \begin{array}{c} n+k-1\\ k-1 \end{array} \right)
\left[ \frac{ \bar{n} / k }{ 1+(\bar{n} / k)} \right]^n
\frac{1}{[1+(\bar{n} / k)]^k} \, .
\label{nbd}
\end{eqnarray}
The parameter $k$ characterizes the width of the distribution. For $k
\rightarrow \infty$ we recover a Poisson distribution, the distribution with
the smallest width. One can see from the figure that at 900 GeV a single
negative binomial (NBD) cannot describe the data anymore.
A double negative binomial (DNBD) of the form
discussed in eq. (\ref{double}) on the other hand describes it very well.
The question remains to what extent a similar analysis might be able to reveal
rare events of quark--gluon plasma production at AGS.
A rough criteria for the observability of such structure in distributions of
the form (\ref{double}) is
\begin{eqnarray}
\frac{2}{\sqrt{N_{\rm bin}}} P_{N_{\rm qg}}^{\rm had} =
q P_{N_{\rm qg}}^{\rm qg} \, .
\label{criteria}
\end{eqnarray}
Here $N_{\rm bin}$ is the number of observed central Au + Au collisions
for which the central multiplicity of the bin is $N_{\rm qg}$.
The right--hand side of eq.
(\ref{criteria}) is the magnitude of the rare events to the overall
multiplicity, while the left hand side gives the statistical resolution.
The assumption here is that $q$ is small, so that at $N_{\rm qg}$ we
can use $P_n \sim P_n^{\rm had}$ for the left--hand side.
To obtain a feeling for the shape and applicability of eqs. (\ref{nbd})
and (\ref{criteria}) we plot in Figures 10 and 11 different negatively
charged particle multiplicity distributions as might be expected for
central Au + Au collisions at AGS with $E_{\rm beam} = 11.6$ GeV/A. From
{}~\cite{multi} we obtained the mean for purely hadronic events to be
$N_{\rm had} = 145$. This is slightly larger than the value
$N_{\rm had} = 131 \pm 21$ cited in \cite{multi} for $355 \pm 7$
participating nucleons since we are assuming that all $2 A$ nucleons
are participating in the collision. The result depicted in Figure 8
for the upper limit on the additional number of negatively charged
pions produced per participating baryon allows us to deduce an upper
limit of $N_{\rm qg} = 193$ on the mean for the events with quark--gluon
plasma production. In Figure 10 we plot the
negatively charged particle multiplicity distribution defined in eq.
(\ref{double}) for different values of the probability $q$. We use Poisson
distributions for $P^{\rm had}$ and $P^{\rm qg}$ and take the upper limit
for rare events $N_{\rm qg} = 193$ as the mean for $P^{\rm qg}$.
A shoulder develops for small $q$ and becomes more pronounced the
larger $q$ is. In Figure 11 we fix $q=0.1$ and investigate the effect
of different values of the mean $N_{\rm qg}$ of the distribution for events
with some quark--gluon plasma production. If this mean is close to the mean
of purely hadronic events we will only find some broadening of the overall
distribution. This would be the case if the phase transition is weakly
first order or second order. For larger $N_{\rm qg}$ we begin to see
a well established shoulder develop. For large $N_{\rm qg}$ a second
maximum appears.
It is clear that the exact values of the probability $q$ and of
the mean $N_{\rm qg}$ of rare events will
be crucial for the experimental observation of a phase transition.
We have provided a first glimpse into this problem, but in the end
it is up to experiment to discover
new physics in multiplicity distributions at the AGS.
\section{Summary and Conclusion}
We have estimated the probability that hard nucleon-nucleon collisions
initiate the formation of seeds of quark-gluon plasma at AGS energies.
Based on our previous studies we know that these will grow rapidly
to convert most of the superheated hadronic matter to quark-gluon plasma.
Our estimates are based on reasonable assumptions and approximations
to the kinetic theory of hadronic physics. Better estimates could
be made using event simulators like RQMD and ARC together with
more detailed knowledge of multi-particle production in nucleon-nucleon
collisions. We find that anywhere from 1\% to 100\% of central
Au + Au collisions should lead to significant quark-gluon plasma
formation. A major assumption is that there is a phase transition
and that it is first order.
We have already proposed that the formation of plasma in rare events
should have an observable consequence for hadron interferometry,
deuteron production, and the meson multiplicity distribution.
In this paper we have studied the effect on the multiplicity
distribution. It would be observable as a shoulder or second
maximum at some multiplicity higher than the most probable one.
If there is a phase transition but it is second order or weakly
first order then the effect will be much more difficult to see.
We eagerly await the results of experiments.
\section*{Acknowledgements}
We thank R. Venugopalan and C. J. Waddington for stimulating discussions
and L. Csernai and P. Lichard for comments on the manuscript.
This work was supported by the U.S. Department of Energy under
grant number DE-FG02-87ER40328.
|
\section{Introduction}
Ly$\alpha$ absorption clouds, observed in the spectra of quasars, are
numerous and detectable to high redshift. As intervening absorption
systems, they are tracers of the evolution of the
gaseous content of the Universe.
By
understanding the extent to which these absorbers are associated with
clusters and large-scale structures of galaxies, we hope to trace the
evolution of these structures as well.
The connection between low redshift Ly$\alpha$ absorbers at z$\leq$0.5
and their high redshift counterparts is presently unclear. The high
redshift Ly$\alpha$ absorbers display very little velocity
correlation. They have been proposed to originate from intergalactic
clouds and, consistent with the absence of clustering, would not be
associated with galaxies \markcite{(Sargent et al. 1980)}.
Groundbased studies of these absorbers have shown strong evolution in
their number density \markcite{(e.g., Bechtold 1994)}. The situation
at low redshifts may be different. The study of Ly$\alpha$ absorbers
at low redshifts has been made possible through the use of the Hubble
Space Telescope (HST) to obtain ultraviolet spectra of moderate
redshift quasars. These observations have revealed a larger number of
absorbers than was initially expected based on the extrapolation of
the evolution observed at higher redshifts \markcite{(Bahcall et
al. 1991; Morris et al. 1991; Bahcall et al. 1993a)}.
They appear to
be more clustered than high redshift absorbers, although to a lesser
degree than galaxies \markcite{(Bahcall et al. 1995)}. The
relationship between Ly$\alpha$ clouds observed at low redshifts and
large-scale galaxy structures can in principle be determined directly
by comparing the observed distribution of the Ly$\alpha$ absorbers
detected in QSO spectra with the distribution of galaxies in the
fields of the same quasars.
There have been several examples of individual Ly$\alpha$ absorption
lines which appear to be associated with galaxies lying at the same
redshift. Such matches have been found in the directions of the
quasars H1821+643 \markcite{(Bahcall et al. 1992)} and PKS 0405$-$123
\markcite{(Spinrad et al. 1993)}
with absorbers having EW $>$ 0.32 \AA.
The matches typically are found to
have impact parameters ranging from 70$h^{-1}$kpc to 160$h^{-1}$kpc
(h=H$_o$/100 km/s/Mpc)
and lie within 500 km/s of the Ly$\alpha$ absorption line redshift.
More recently, Lanzetta et al. (1995) have surveyed galaxy redshifts
in the fields of several quasars observed by HST and found 11
galaxy/absorption system matches with impact parameters
$\lesssim$160$h^{-1}$kpc. They conclude that at least 0.32 $\pm$ 0.10
Ly$\alpha$ absorption lines are due directly to intervening galaxies
having large halo regions of hydrogen gas.
While such studies will
help determine if some individual galaxies are associated with Ly
$\alpha$ absorbers, they do not allow us to investigate the relationship
between the absorbers and the large-scale structures of galaxies. These
studies lack complete redshift sampling of the galaxies in the
individual quasar fields and cover limited regions (in
angular extent) around each quasar. As a result, the complete
line-of-sight distribution of galaxies is not well determined on size
scales comparable to that of
clusters of galaxies and it
is not generally possible to determine if the observed galaxies are
part of larger structures (groups or clusters of galaxies).
There is evidence that Ly$\alpha$ clouds could be distributed
randomly with respect to regions of high
galaxy density. In the direction of the quasar 3C273, one of the
strongest Ly$\alpha$ lines is found in a region where the nearest
luminous galaxy is 10 Mpc away \markcite{(Morris et al. 1993)}. The
Morris et al. study also found that in a region of high galaxy
concentration in the same field, no Ly$\alpha$ lines are found within
36 Mpc. Stocke et al. (1995) have made
similar findings using the CfA
galaxy redshift survey in the direction of Mrk 501. They find no
galaxies brighter than M$_{B}$=-16 within 100$h{_{75}}{^{-1}}$ kpc
of a $>$ 4$\sigma$ Ly$\alpha$ absorption line detection.
Although the galaxy coverage is very good for these studies, they
probe a relatively small path length in redshift space and therefore
intercept fewer strong Ly$\alpha$ lines (observed equivalent widths
greater than 0.32\AA) that can then be compared with the galaxy
distribution. In general, all the Ly$\alpha$ lines studied in the
Stocke et al. work are weaker than the typical Ly$\alpha$ lines
observed and studied at higher redshift.
In this paper we investigate what data are necessary to adequately
determine the extent and nature of the association between the
large-scale galaxy structures and Ly$\alpha$ absorbers observed along
the sightlines to distant quasars. Obviously, catalogues of both
absorbers and the large-scale structures are needed. Not as clear is
the exact nature and size that these databases must take. In the case
of the absorption line catalogue we will assume that for the near
future the largest and most complete catalogue available will be that
being compiled by the HST Quasar Absorption Line Key Project (Bahcall
et al. 1993a; Bahcall et al. 1995). Their observations will provide a
large and homogeneous catalogue of strong Ly$\alpha$ absorption
lines toward over 80 quasars with redshifts between 0.15 and 1.9.
However, the number of Ly$\alpha$ absorbers along each individual
line-of-sight is greatly reduced from what we see at high redshift and
the entire redshift path length for each line-of-sight is not observed
for every quasar (in fact most of the quasars were not observed below
1600 \AA). Given this catalogue of Ly$\alpha$ absorbers, we have
conducted various simulations designed to determine what is required
of a redshift survey in the fields of these quasars in order to test
the following hypotheses: 1) low redshift Ly$\alpha$ absorbers are
uncorrelated with the large-scale structures of galaxies traced by the
peaks in the galaxy distribution; 2) some fraction of the Ly$\alpha$
absorbers is associated with the peaks in the galaxy distribution with
the remaining absorbers being unassociated. Specifically, we
determine the characteristics and number of the galaxies that need to
have measured redshifts along each sightline and the total
number of QSO sightlines that must be studied in order either to
falsify the first hypothesis or to determine the fraction of associated
absorbers if the second hypothesis is valid.
\section{Analysis}
Several factors make the investigation of the relationship between
strong absorbers and large-scale
structures difficult at redshifts less than
0.2. First, there are extremely few strong Ly$\alpha$ absorption
systems at very low redshift (intrinsically, the volume density is low
as a practical consequence of the limited amount of path length
observed at redshifts less than 0.2). Second, the volume of space
surveyed at low redshift is relatively small, yielding few large-scale
structures for comparison. Third, those structures which are present
are difficult to identify without a very wide field (on the order of a
one degree diameter) redshift survey. As a result, most of the
information in studying the relationship between absorbers and large
scale structures (at least for the strong lines contained in the Key
Project database) will come by comparing the cluster and group
distribution to the absorbers at redshifts between 0.2 and 0.5.
In this section we try to determine the rough characteristics (e.g. number
of galaxies in each field, how bright, over what redshift range, for
how many quasar fields) an incomplete redshift survey would need in
order to test either of the hypotheses stated at the end of the
introduction.
\subsection{Sampling the Galaxy Distribution}
If only a small number of galaxy redshifts are obtained
near the sightline of a quasar, the redshift distribution of
these galaxies does not show obvious peaks and voids and therefore does
not yield much information about the location in redshift space of
large-scale structures of galaxies.
We are interested in determining
the minimum number of galaxies which would allow us to best estimate
the location of large-scale structures in redshift space.
In this section, we determine the minimum number of
galaxies which must be observed near the quasar line-of-sight
satisfying the following
criteria: 1) the subset has the largest possible fraction of its
galaxies in the
``peak'' regions in redshift space and 2) every peak in the true galaxy
distribution is represented by at least one galaxy in the subset.
To model the effects of incomplete sampling of the true distribution
of galaxies we simulated limited observations of ``true''
distributions as defined from two very extensive galaxy redshift
surveys along different sightlines. The components of these surveys
are described in Tables 1 and 2 \markcite{(Peterson et al. 1986;
Broadhurst et al. 1993; Colless et al. 1990; Broadhurst 1994)}. One
survey is in the direction 1043+00 and contains a total of 213 galaxy
redshifts. The pencil beam diameter of the survey is about
$\sim$32$\arcmin$ corresponding to 6.8$h^{-1}$Mpc at z=0.25. The
survey samples galaxies down to an absolute magnitude of
M$_{B}\simeq$-18 out to z$\simeq$0.54 where we have
assumed H$_o$ = 75 km/s/Mpc.
The second survey is in the
direction of the South Galactic Pole at 0055-28 and contains 161
galaxy redshifts. The diameter of the survey beam is about
$\sim$21$\arcmin$ corresponding to 4.6$h^{-1}$Mpc at z=0.25. This
survey samples galaxies down to an absolute magnitude of
M$_{B}\simeq$-18 out to z$\simeq$0.46. Both surveys include a much
broader cone covering about 3$\deg$, but extending to redshifts
of only $\sim$ 0.1. The nominal completeness of the various components of each
survey is listed in the tables. These surveys provide us with an
empirical test bed for modeling the limitations of incomplete redshift
surveys in representing the large-scale structures contained in the
survey volume.
Figures 1a and b show the redshift distributions of these
two surveys collapsed along a single line-of-sight in redshift space.
There are obvious peaks in the distributions noted by Broadhurst et
al. (1990) as the pencil-beam intersects large-scale structures.
The survey pencil-beam diameters are close to optimal for
detecting wall-like topologies on scales comparable to those revealed
in the CfA surveys \markcite{(Szalay et al. 1991)}.
Since we
are interested in the large-scale distribution of galaxies,
we have conducted our tests using the galaxy distribution in redshift
space alone without considering the effects of different impact
parameters of the individual galaxies to the QSO sightline.
Over the
angular fields we are considering, the distribution of the galaxies in
redshift space provides ample information about the locations of
groups and clusters of galaxies along a sightline. Our tests were
designed to simulate actual ``observations'' of galaxies along these
lines-of-sight through randomly selected subsamples of the total data
set representing the objects whose redshifts are obtained in
a given limited redshift survey.
Using these surveys as a representation of the ``true'' universe, we
will investigate how well selected subsets of the sample represent
the large-scale distribution of the galaxies including the incidence
of clusters or peaks in the redshift distribution.
Our first step is to define a weighting function in redshift space
that indicates association with peaks in the redshift surveys.
For the redshift distributions of Figures 1a and b,
we have chosen a histogram representation with
bin size of $\Delta$z = 0.01 so that a typical galaxy cluster
or group in the data would be sampled by at least 2 bins.
The location and width of the peaks were
determined mainly through visual inspection
based on apparent over-densities in the galaxy distribution
and the width of those over-densities at half of their maximum height
in the histogram.
The peak locations and widths can also be identified by determining
the local noise level (standard deviation)
within a 3 bin radius of the assumed peak.
Gaussian statistics were used to determine the standard deviation from 6
bins consisting of 3 bins on either side of the peak, excluding those
associated with nearby peaks.
Our ``peaks'' are those bins which are
$\ge$3$\sigma$ above this noise level.
Once the location and width of peaks in each galaxy distribution
has been determined, we can define our weighting function.
The weighting function is designed so that regions in redshift space that
are ``associated'' with a peak have a weight of 1.0 and regions that are
well outside the peaks have weights of 0. We can consider each peak
as a Gaussian shape having a full-width at half-maximum equal to the
peak width. The ``peak'' regions in our
weighting function are then defined as the peak center $\pm$ 1$\sigma$
and the function value within these regions is 1.0. Beyond $\pm$
1$\sigma$ the function value behaves as a Gaussian, trailing off towards
0 in the ``void'' regions of redshift space (see Figures 2a and b). In
this way, each peak has a finite width in the weighting function.
The weighting function can be used as a measure of the degree of
concentration of any subsample of galaxies to the peaks.
To do so, the galaxies in a subsample are assigned an initial delta function
of unit amplitude, then weighted by multiplication with the galaxy
distribution functions in Figures 2a and b for each
of the two redshift surveys respectively.
The mean function value for the weighted sample measures the concentration
in peaks of the redshift distribution, or the averaged probability that
an individual galaxy in the sample has a redshift associated with a peak.
Obviously, not all galaxies in
the redshift surveys fall within a peak according to the weighting function
of Figure 2.
To satisfy the first criterion mentioned at the beginning
of this section, we investigated what constraints could be placed
on subsample selection to allow a higher fraction of the
galaxies to fall within the ``peaks''.
These constraints will point to a sampling strategy for determining
the redshift peaks in newly observed samples with the best attainable
reliability.
Figures 1a and b show the galaxy distributions with
all of the galaxies from each of the smaller surveys providing us with
our ``true'' map of the universe. The low z peaks are more
easily identified
due to the inclusion of the 3$\deg$ diameter
field which
extends to only z$\lesssim$0.1. However, the deep portion of these
surveys, necessary in order to consider comparison
between the absorbers and large-scale structures, only covers the
inner 20$\arcmin$ to 30$\arcmin$ of each survey. The hatched region
in Figures 1a and b represents the galaxy distribution within this
smaller cone. When only the smaller angular field is considered, it
becomes difficult to identify the low redshift peak in the
distribution.
Since the data set we are using to represent
the ``true'' universe does not allow us to define a peak at
low redshift, and for the reasons described at the beginning of this
section, we place a lower limit of z=0.2 for
identifying peaks in cones of 30$\arcmin$ or less.
An upper limit at z=0.4 is imposed by the
limitations for obtaining redshifts for a reasonable sampling of the
galaxy luminosity function with 4-m class
telescopes. Therefore, we have little information from the two
surveys used here about the true galaxy distribution above z=0.4 and
cannot determine how well galaxies beyond this redshift represent
the peaks and voids of the distribution.
Intrinsically bright galaxies have a somewhat higher probability
of being in the
peaks of the distribution.
Supporting evidence is found in the fact that eliminating the data
from the redshift surveys having z $<$ 0.2
increases the mean function value determined from the remaining
galaxies for both redshift surveys.
The less luminous objects
observable at low redshifts, seem to be more uniformly distributed.
Limiting the observed
galaxy absolute magnitude range for a random sampling of the galaxies
in the field improves the probability that a given galaxy is a member
of a peak in the true galaxy distribution. The redshift surveys we
are using have galaxy absolute magnitudes ranging from M$_B\simeq$ -18
to M$_B\simeq$ -21 based on their apparent magnitudes with H$_o$=75
km/s/Mpc. If we exclude the few galaxies which are fainter than M$_B$
= -18, the concentration in peaks is increased. While raising
this lower limit may increase the concentration further it also greatly
decreases the number of observable galaxies, i.e. there aren't enough
luminous galaxies available to locate and define the peaks in the
galaxy distribution. We therefore have chosen to constrain the range
of galaxy absolute magnitude to M$_B$ $\leq$ -18 which is $\sim$
0.44$L^\star$ \markcite{(Marzke et al. 1994)}. This corresponds to an
apparent magnitude of $B$ $\leq$ 23 at z=0.4. Eliminating galaxies
with M$_B$ $>$ -18 increases the mean function value from the
remaining galaxies in each of the two redshift surveys.
Once we have
maximized the fraction of galaxies falling within the peaks
in an ``observed'' subset,
we can determine the minimum size subsample that retains
this same fraction of galaxies associated with peaks in
the galaxy distribution.
In addition, these subsets must also have
at least one galaxy in the redshift range of each peak in the
true galaxy distribution. In this way, we can make certain that
all peaks in the true galaxy distribution are represented.
To do this,
we simulated ``observations'' of the true universe by randomly selecting
galaxies in the test surveys
through sub-cones 5$\arcmin$ in diameter using numerous cones to span
each survey out to the largest effective angular size of the
survey. For our purposes, simulating observations of galaxies at
significant redshift (out to z=0.4), the ``effective angular size'' is
constrained to be the largest angle in each test survey for which
galaxy redshifts out to at least z=0.4 are available. For this reason
we limit our simulations to the the inner 35$\arcmin$ x 32$\arcmin$
of the 1043+00
survey and the inner 23$\arcmin$ x 10$\arcmin$ of the
South Galactic Pole survey.
This same simulated observational procedure was repeated with
increasing subcone sizes until
the maximum size of the survey was reached.
For each ``observation'', the galaxies found within the cone were
assigned an intial delta function of unit amplitude, then weighted by
multiplication with the galaxy distribution function in Figures 2a and b.
By averaging the weighted amplitudes of
all of the galaxies observed within a cone, a mean function value is
determined. For example, if half of the galaxies in a cone fall
within 1$\sigma$ of the central redshift of a peak and the other half
fall completely outside, the mean function value for that cone is 0.5,
indicating a 50\% chance that a given galaxy in that subset was
selected from a peak in the true galaxy distribution.
Figures 3a and b show the number of galaxies observed in a cone vs.
the mean function value for each of the two redshift surveys with the
data sampling limits as described above. Each dot represents a
different sub-cone of the total survey within which $n$ galaxies have
been observed. These plots contain cone sizes from 5$\arcmin$ in
diameter to the largest angular size possible within the survey
limits; 35$\arcmin$ x 32$\arcmin$ for the 1043+00 survey and
23$\arcmin$ by 10$\arcmin$ for the SGP survey. It is clear that if
all of the galaxies in these surveys which are between the stated
redshift and magnitude limits are observed, then the mean function
value is 0.68 for the 1043+00 survey and 0.74 for the South Galactic
Pole survey. As we examine galaxy sets containing fewer and fewer
observed galaxies, the typical deviation from the mean function value
for a given subset becomes
greater than $\sim$5\% for subsets containing fewer than
$\sim$18 galaxies. For samples with sizes below this limit,
the mean function value, expressing the probability that a given galaxy in
that subset lies within 1$\sigma$ of a peak in the true distribution,
becomes very uncertain.
These simulations of galaxy observations along a specific
line-of-sight suggest that the probability of observed galaxies lying
in redshift peaks reaches a maximum between 0.68 and 0.74 where we
have limited our redshift range to 0.2$\leq$z$\leq$0.4 and
placed a lower limit on the absolute magnitude of M$_B\leq$ -18
corresponding to an apparent magnitude of $B\leq$ 23 at z=0.4.
A subsample of
at least $\sim$18 galaxy redshifts are needed for a representative
sample where
$\sim$70\% of the galaxies fall within 1$\sigma$ of a peak in
the true galaxy distribution. The angular field of view necessary to
obtain this minimum number of galaxy redshifts is r $\simeq$
11$\arcmin$ in the 1043+00 survey and r $\simeq$ 7$\arcmin$ in the
South Galactic Pole survey. These spatial ranges correspond to
sampling regions of space $\sim$2.2 Mpc in size at
z=0.3.
The angular diameters quoted here, however, are dependent on the
completeness and efficiency for the redshift surveys we used.
The selection efficiency for obtaining galaxy redshifts is critical
in determining the angular diameter of the cone required to
measure a sufficient number of objects. A simple integration of the
galaxy luminosity function
\markcite{(Marzke et al. 1994)}, in a truncated cone of r = 10$\arcmin$
and redshift limits of 0.2 to 0.4 suggests that we would find $\sim$120
galaxies mith M$_B\leq$ -18, about six times the number of galaxies
within the same constraints for the surveys used here.
It is therefore possible to limit the
angular area required for search around each quasar by increasing the
efficiency with which redshifts are obtained. There is a lower limit
set by the physical area subtended by the large-scale structures
themselves. One might consider a Mpc or so as the minimum diameter
below which the search becomes more relevant to individual objects
rather than clusters or associations. A field with a radius of
4$\arcmin$ at z=0.2 would encompass over 1 Mpc
and provide
enough galaxies for redshift measurement to meet our criterion
given a $\sim$100\% efficiency for obtaining redshifts. With
a $\sim$60\% efficiency, the minimum number of galaxy redshifts could be
obtained within a 5$\arcmin$ radius field.
There are $\sim$10-12 sets containing 18 or more galaxies in each
of the two redshift surveys. These subsamples
can then be binned in redshift space in the same manner as the entire
redshift survey. If we consider each galaxy in these subsets as a
peak, we find that each peak, defined from the total galaxy
distribution within the 0.2$\leq$z$\leq$0.4 redshift range, is
represented. By assuming that each galaxy represents a peak in the
galaxy distribution, all true peaks are located. However, only
$\sim$70\% of the galaxies are associated with true peaks;
approximately 30\% of the galaxies will actually lie in the ``void''
regions of the true galaxy distribution. We find that overestimating
the number of peaks by 30\% is
the minimum error which can be attained in defining
peaks in redshift space while also minimizing the total
number of galaxy redshifts obtained.
\subsection{Comparing the Galaxy and Lyman Alpha Cloud Distributions}
Comparison of the distributions of galaxies and absorbers requires not
only the sample of galaxies, for which the determination of
large-scale structure was discussed in the previous section, but also
a sample of absorbers. For this paper we will assume that low
redshift Ly$\alpha$ absorption systems detected on lines-of-sight to
quasars at redshifts greater than 0.4 are consistent in distribution
and number with what has been observed by the HST Quasar Absorption
Line Survey (Bahcall et al. 1993a). Therefore, determining the number
of Ly$\alpha$ clouds necessary to test the two hypotheses stated in
the introduction can be restated as determining the minimum number of
lines-of-sight that need to be observed for absorbers and galaxies.
We are assuming that the absorption lines found along
a line-of-sight will be drawn from the simplified line distribution
function
\begin{equation}
{{dN}\over{dz}}={({{dN}\over{dz}})}_o{(1+z)}^\gamma
\end{equation}
with (dN/dz)$_o = 18$ and $\gamma = 0.3$ as determined from HST
observations by Bahcall et al. (1993a) based on the detection
of Ly$\alpha$ lines having rest equivalent widths of 0.32\AA
or greater.
To model the case of no association between large-scale structure
and absorbers, simulated samples of Ly$\alpha$
absorbers were generated with no velocity correlations
on small scales and with a line-of-sight
density evolution with redshift as defined above.
Eq. 1 was used to determine the probability that a line would exist
($\Delta$N) within a certain redshift bin ($\Delta$z). The redshift
bin size was chosen to be the
resolution element size for the
HST Quasar Absorption Line Survey which is $\Delta$v=270 km/s.
We then divided the redshift space between 0.2 and 0.4 into bins
of this size.
A random number generator was used to produce a number between 0 and
1 for each bin. If that number was less than the probability
$\Delta$N determined for that bin, a Ly$\alpha$ line would be generated at
that redshift.
This same procedure was repeated to simulate line lists from many
lines-of-sight with the total distribution with redshift consistent with
the global distribution found by Bahcall et al.
To obtain a quantitative measure of association between the peaks
in the galaxy distribution and absorbers,
the redshift distribution of each random absorption
line list was then weighted by the
galaxy redshift distribution function (see Figures 2a and b)\footnote
{All tests comparing the Ly$\alpha$ line distributions to that of
galaxies along a sightline were performed separately using both redshift
surveys. Because the results from each survey were consistent with
one another, we will only present results
for the 1043+00 survey in the text and figures.}.
We compare the distribution of Ly$\alpha$ line
function values to similarly computed distributions of function values
for subsets of galaxy redshifts drawn from the 1043+00 and SGP
surveys.
The number of galaxies along each sightline was chosen in a manner
consistent with the discussion in section 2.1, providing the minimum
number of galaxy redshifts to represent the peaks in the galaxy
distribution. The dotted line in Figure 1 shows the arbitrarily
normalized galaxy selection function for the 0.2$\leq$z$\leq$0.4
range for each survey based on a Schechter luminosity function
(Marzke at al 1994). The effects of this function are ignored in
our simulations when choosing galaxy subsets for comparison with
the Ly$\alpha$ line lists
since variations of this function are small over this redshift range.
As the number of sightlines
increases, the number of galaxies in the comparison sample
must also increase. Since the previous experiment indicates that
$\sim$18 galaxies are necessary to characterize adequately the galaxy
distribution along a single sight line,
we conducted this test assuming that the minimum number
of galaxy redshifts will be measured for each sightline.
A set of 18 galaxies is compared with the Ly$\alpha$ line list for a
single sight line,
a set containing 36 galaxies is compared with 2 Ly$\alpha$ line lists, etc.
Since the galaxies in these distributions are drawn from the redshift
range of z=0.2 to 0.4, the Ly$\alpha$ line lists generated contain
lines within this range which results in an average of 3.9 absorbers
per line list.
The Kolmogorov-Smirnov test (KS test) was used to find the level of
confidence at which the distribution of weighting function values for observed
galaxies is different from that of the locally Poissonian distribution
of Ly$\alpha$ lines. Figure 4 shows this KS
probability as the number of observed sightlines increases. To show
that the two distributions are different at the 90\% confidence level,
at least 6 lines-of-sight must be observed or $\sim$24 Ly$\alpha$
lines in total.
At least 8 lines-of-sight are
necessary for 95\% confidence and 12 are required to show that they
are different at the 99\% confidence level.
This exercise demonstrates that the distribution of
Ly$\alpha$ lines in redshift space, constructed to
show locally Poissonian statistics,
obviously differs from a representative distribution of galaxies which
shows clustering. If the Ly$\alpha$ lines are distributed in this
way, we would need to study only $\sim$8 to 12 sightlines for
absorbers (observed at the level of the HST Key Project observations)
and galaxies (with a limited redshift survey as discussed above) in
order to recognize that the absorbers and large-scale structures of
galaxies are not distributed in the same manner.
Next we explored the possibility that only a fraction of Ly$\alpha$
lines are randomly distributed while the remainder are associated with
the peaks in the galaxy distribution. We
compared the degree of concentration in the peaks for
Ly$\alpha$ line lists generated from the power-law equation
to that for Ly$\alpha$
lines distributed like the peaks in the galaxy distribution.
Remember, however, that unless we have a very large number of galaxy
redshifts along the quasar sightline, we cannot know the location of
actual peaks in the galaxy distribution with absolute certainty. Our
earlier simulations indicate that with subsets of at least $\sim$18 galaxies
(meeting the selection criteria detailed above and assuming each
of these galaxies is located in a ``peak'') we overestimate the number
of peaks by 30\%.
The weighting function values (which measure the degree of association with
peaks) for these generated Ly$\alpha$ lines
must therefore be corrected for a 30\% excess in apparent
associations. We chose to do that in the simulations by generating an
``associated'' line list from the ``observed'' galaxy sample, then
applying the ``true'' weighting function to assign zero weight
to the 30\% which are spurious associations.
The peak associated Ly$\alpha$ lines are generated in much the same way
as those with global properties characterized by Eq. 1 as described earlier
in this section. The only difference is that we have substituted the
probability function of Eq. 1 with
an empirically determined probability function representing
the peaks. In other words, rather than using the pure power-law
equation for dN/dz in Eq. 1 to determine the probability that a line
will be placed in a particular $\Delta$z redshift bin, we are using the
global trend for dN/dz modified by the location and width of
peaks as determined
by subsets of galaxies containing at least $\sim$18 galaxies.
This distribution assumes the same number density of low-redshift
Ly$\alpha$ lines as described by Eq. 1 but
redistributed on small scales to correlate with the peaks.
The Monte-Carlo generated Ly$\alpha$ lines were then weighted by the
true galaxy distribution functions in Figures 2a and b so that each line
received a value based on its redshift location with respect to true
peaks in the galaxy distribution. Figure 5 shows the normalized
cumulative distributions of these function values for 10000 Ly$\alpha$
lines distributed with no association to the peaks in the galaxy
distribution (solid line) and 10000 Ly$\alpha$ lines distributed like
the peaks in the galaxy distribution (dashed line).
Within the distribution of random Ly$\alpha$ lines there are many that
fall at or
near a function value of 0 since many random lines will fall
between peaks based on our weigting function in Figures
2a and b. For the peak-distributed lines (dashed line), note that
fewer fall at 0 and many more have function values of 1.0 since they
have been generated to be associated with peaks in the galaxy distribution.
Still, some have function values at or near 0 due to the fact that the
peak-associated lines have been generated based on peak locations as
defined from subsets of $\sim$18 galaxies. Since only 70\% of these
galaxies actually fall in a true peak, we have generated some
peak-associated lines at redshifts where an actual peak doesn't
exist. These lines will have values at or near 0 since they don't fall
in the ``true'' peak regions in redshift space.
Any mixture of the two distributions in Figure 5 will
produce a cumulative distribution located between them determined by
the percentages of each parent population contained within it.
The aim of our test is to characterize to what accuracy we can
determine the composition of an observed distribution of Ly$\alpha$ lines
taken from several lines-of-sight based on the proximity of each
Ly$\alpha$ line to peaks in the galaxy distribution. In other words,
if we obtain a sample of Ly$\alpha$ line redshifts, how accurately can
we determine the fractional contribution of each of the two parent
distributions, purely unclustered lines and galaxy peak associated
lines, and what is the minimum number of Ly$\alpha$ lines which must
be observed to make this determination?
Our approach differs from that taken by
Bahcall et al (1993a) who calculated the number of Ly$\alpha$ lines
necessary to detect clustering within the Ly$\alpha$ sample alone.
We are interested in determining the number of Ly$\alpha$ lines needed
to identify a population clustered with galaxies
when we are able to make use of additional information about the
distribution of the galaxies along the same sight lines as
the absorbers.
For our simulation the
distribution of galaxies
is determined from the simulated limited
redshift surveys described in section 2.1.
Our approach is to generate samples of lines for which the fractional
contribution of each parent distribution is known. The sample is then
compared through a KS test to a set of distributions with a range in
fractional composition of the two parents
(Figure 5). This computation produces a distribution of KS probability
values peaking at the fractional mixture of the two parents which best
fits the sample. Since the KS test determines the probability that
two distributions are different, a large KS probability indicates a
lack of difference or a ``best fit''.
Many random draws of samples containing the same number of lines, for
example, in a 50:50 ratio will produce a range of KS probability
distributions where the best fit will vary around $\sim$50\%. To take
into account the variation in peak location and width of the KS
probability distribution, we generated distributions for 100 random
draws for each sample of Ly$\alpha$ lines and determined the mean
fractional composition value and its variance at the KS probability
peak. The range in fractional composition of parent populations which
includes 95\% of the random draws is represented by the mean peak
value $\pm$ 2 times the standard deviation of the distribution. We
computed this range for samples containing various total numbers of
Ly$\alpha$ lines and fractional compositions. We find that the
standard deviation for Ly$\alpha$ line sets containing the same total
number of lines remained roughly the same regardless of composition
value.
Figure 6 reveals that as more Ly$\alpha$ lines are considered, the
range of fractional composition values decreases allowing for the true
composition to be more accurately determined. At 150 Ly$\alpha$
lines, approximately 38 lines-of-sight, enough Ly$\alpha$ absorption
lines are observed within the 0.2$\leq$z$\leq$0.4 redshift
range to determine the fractional composition to within 10\% of the
true value for $\sim$95\% of the generated samples. The accuracy
improves slowly as the number of lines-of-sight increases. If fewer
than 5 sightlines are observed, these tests suggest that it is
impossible to determine the composition of the sample, i.e what fraction
are associated with large-scale structures if some fraction are random
with respect to these structures. 5 sightlines are equivalent
to only $\sim$20 Ly$\alpha$ lines in the redshift range being considered.
We find that
the measurement of Ly$\alpha$ absorbers from at least 38 sightlines is
necessary (equivalent to 150 Ly$\alpha$ lines), in addition to the
minimum number of galaxy redshifts
needed, to show that the population of Ly$\alpha$ absorbers is
composed of a mixture of those which are distributed like the peaks in
the galaxy distribution and those which are uncorrelated with the
peaks in the galaxy distribution.
\section{Discussion}
Recent surveys to identify and obtain redshifts
for galaxies projected near quasar
sightlines have been designed to study the relationship between absorbers and
individual galaxies. Other surveys have concentrated on the association of
quasars with host galaxy clusters. Often the field of view is limited by the
cassegrain spectrograph with multi-object capability. Such samples can serve
as the starting point for defining the larger-scale structures. They may not
be adequate to do so in their current form; they were not designed to be.
Two examples from current surveys illustrate the point.
The quasar PKS 0405-123 has been observed with HST revealing 14 Ly$\alpha$
absorbers within the redshift range of z=0.081 to 0.540
\markcite{(Bahcall et al. 1993b)}. Recently, Ellingson et al. (1994)
published a list of 29 galaxy redshifts in this field ranging from
z=0.16 to 0.66. This redshift survey covers a field 5.9$\arcmin$ by
3.8$\arcmin$ around the quasar and is 78\% complete to r=21.5. The
redshift range of this survey is comparable to the surveys used in our
significance tests. In the optimal redshift range determined for the
surveys used in our simulations (z=0.2 to 0.4), 10 galaxy redshifts
are measured which meet our absolute magnitude criterion of
M$_B\leq$-18. We assume B-R colors for these galaxies of up to 1.75
(Colless et al. 1990). Figure 3 indicates that a sample of 10
galaxy redshifts does not meet the minimum sample size requirements to
ensure that 70\% of the galaxies will
fall within 1$\sigma$ of a peak in the
galaxy redshift distribution.
As another example, consider the available observations of the field
of the quasar 3C~351.
This quasar has also been observed with HST revealing $\sim$16
Ly$\alpha$ lines within the redshift range of z=0.092 to 0.370
\markcite{(Bahcall et al. 1993a)}. Included in the survey
of Lanzetta et al. (1995) are redshifts for 10 galaxies in this field,
ranging from z=0.07 to 0.370. Their survey of this field covers
$\sim$ 5$\arcmin$ in diameter and is 57\% complete down to r=21.5.
Within their sample, 4 galaxies meet our redshift range and absolute
magnitude criteria. Again, the small sample size does not ensure that
70\% of these galaxies
lie within 1$\sigma$ of a large-scale
structure.
More extensive redshift coverage of galaxies in wider fields around
these quasars would allow for a better determination of the
large-scale distribution of galaxies along the sightline. Although
some observations of galaxies in the fields of many more of the HST
observed quasars exist, not enough redshifts have been measured to
yield statistically significant results. Ideally, it would be
necessary to obtain enough galaxy redshifts along the line-of-sight to
a quasar having a redshift at or beyond z=0.4 to satisfy the
requirements determined in section 2.1. In this way, all the peaks in the
galaxy distribution out to the QSO redshift would be represented with
an overestimation of 30\%.
We would then need to obtain these surveys in
the fields of at least $\sim$8 QSO's for which HST spectra are
available in order to test the hypothesis that the Ly$\alpha$
absorption clouds are uncorrelated with the peaks in the galaxy
distribution. It would be necessary to obtain these surveys in the
fields of $\sim$38 QSO's with HST spectra to determine
to 10\% accuracty what
fraction of the Ly$\alpha$ absorbers are associated with the peaks in
the galaxy distribution.
\section{Conclusions}
We have conducted various numerical experiments designed to
investigate the significance with which the association between
Ly$\alpha$ absorption clouds and the large-scale distribution of
galaxies can be determined. We have found that in pencil-beam
redshift surveys extending to redshifts of z$\sim$0.5, the maximum
probability of selecting a subsample of galaxies such that every peak
in the distribution is represented by at least one galaxy occurs when
redshifts for at least 18 galaxies are obtained between
z=0.2 and 0.4 with M$_B\leq$=-18 and drawn from an angular radius of
$\sim$10$\arcmin$ around the quasar line-of-sight.
The limited cone size
of the surveys used in these simulations is the main factor pushing us
to the z$\geq$0.2 region to sample volumes of space large enough to
detect large-scale structures. Based upon these redshift surveys, it
would be necessary to have at least $\sim$18 galaxy redshifts in each
quasar field to populate all the peaks in the galaxy distribution
along the line-of-sight. Without knowing the true galaxy distribution
along a sightline, we conclude that $\sim$70\% of the $\sim$18 galaxy
redshifts measured will fall within $\sim$1$\sigma$ of a true peak in the
galaxy distribution.
A typical sightline must be surveyed in a cone of radius
$\sim$10$\arcmin$ to get this number of redshifts down to $B$=23 at
z=0.4 for the surveys used in this study, typical of a 4-meter class
telescope redshift survey.
If the Ly$\alpha$ absorption clouds are uncorrelated with the peaks in
the galaxy distribution, we find that at least 8 lines-of-sight must
be observed to show that the distribution of galaxies and that of the
absorbers is different at the 95\% significance level. However, if
some fraction of the Ly$\alpha$ absorbers is distributed like the
peaks in the galaxy distribution and some fraction is uncorrelated, we
find that $\sim$38 lines-of-sight must be observed to determine the
fraction (to 10\% accuracy) of absorbers which are distributed like
the galaxies.
Our test results clearly indicate that more data are needed in order
to draw reliable conclusions about the extent and nature of the
association between Ly$\alpha$ absorption clouds and the peaks in the
galaxy distribution along the line-of-sight. Fortunately, several
research groups are actively obtaining redshifts for galaxies in the
fields of quasars observed with HST.
\acknowledgments
We would like to thank Tom Broadhurst for providing the redshift
surveys used in this paper in electronic form. Thanks also to Gary
Schmidt, Joe Shields, Jill Bechtold and Ata Sarajedini for helpful
conversations. We thank Simon Morris for his valuable comments
and suggestions for improving this paper. B.T.J. acknowledges
support for this work by NASA
through grant number HF-1045.02-93A from the Space Telescope Science
Institute, which is operated by the Association of Universities for
Research in Astronomy, Incorporated, under NASA contract NAS5-2655.
\clearpage
\begin{planotable}{cccccccc}
\tablewidth{33pc}
\tablecaption{Redshift Survey Data for 1043+00 Field}
\tablehead{
\colhead{\# of Galaxies} &
\colhead{z range} & \colhead{$B$ range} &
\colhead{field size} & \colhead{completeness} &
\colhead{ref.}}
\startdata
70& 0.0 to 0.097& 13.95 to 17.19& 209$\arcmin$.3 x 213$\arcmin$.2& 74\%& 1 \nl
53& 0.0 to 0.213& 17.32 to 19.69& 34$\arcmin$.3 x 33$\arcmin$.5& $\sim$80\%& 4
\nl
108& 0.0 to 0.438& 19.70 to 20.77& 36$\arcmin$.8 x 31$\arcmin$.8&
$\sim$80\%& 4 \nl
32& 0.0 to 0.543& 21.05 to 22.50& 5$\arcmin$.3 x 12$\arcmin$.0& 81\%& 3 \nl
\end{planotable}
\begin{planotable}{lrrrrcrrrrr}
\tablewidth{33pc}
\tablecaption{Redshift Survey Data for South Galactic Pole Field}
\tablehead{
\colhead{\# of Galaxies} &
\colhead{z range} & \colhead{$B$ range} &
\colhead{field size} & \colhead{completeness} &
\colhead{ref.}}
\startdata
75& 0.0 to 0.133& 13.82 to 17.50& 208$\arcmin$.8 x 209$\arcmin$.0& 52\%& 1 \nl
59& 0.0 to 0.444& 20.50 to 21.50& 22$\arcmin$.8 x 9$\arcmin$.5& 84\%& 2 \nl
27& 0.0 to 0.564& 20.71 to 22.42& 5$\arcmin$.3 x 12$\arcmin$.0& 80\%& 3 \nl
\tablerefs{
(1) Peterson et al. 1986; (2) Broadhurst, Ellis \& Shanks 1993;
(3) Colless et al. 1990;
(4) Broadhurst 1994.}
\end{planotable}
\clearpage
|
\section{~~~Introduction}
\def9.\arabic{equation}{1.\arabic{equation}}
\setcounter{equation}{0}
Observational cosmology is entering a new era where it is becoming
possible to make detailed quantitative tests of models of the early
universe for the first time. Such observations are presently the most
plausible route towards learning some of the details of physics at
extremely high energies, and the possibility of testing some of the
speculative ideas of recent years has generated much excitement.
One of the most important paradigms in early universe cosmology is
that of cosmological inflation, which postulates a period of
accelerated expansion in the universe's distant past (Starobinsky,
1980; Guth, 1981; Sato, 1981; Albrecht and Steinhardt, 1982; Hawking and
Moss, 1982; Linde, 1982a, 1983). Although originally introduced as a
possible solution to a host of cosmological conundrums such as the
horizon, flatness and monopole problems, by far the most useful
property of inflation is that it generates spectra of both density
perturbations (Guth and Pi, 1982; Hawking, 1982; Linde, 1982b; Starobinsky,
1982; Bardeen, Steinhardt, and Turner, 1983) and gravitational waves
(Starobinsky, 1979; Abbott and Wise, 1984a). These extend from extremely
short scales to scales considerably in excess of the size of the
observable universe. During inflation the scale factor grows
quasi-exponentially, while the Hubble radius remains almost constant.
Consequently the wavelength of a quantum fluctuation -- either in the
scalar field whose potential energy drives inflation or in the
graviton field -- soon exceeds the Hubble radius. The amplitude of
the fluctuation therefore becomes `frozen'. Once inflation has ended,
however, the Hubble radius increases faster than the scale factor, so
the fluctuations eventually reenter the Hubble radius during the
radiation- or matter-dominated eras. The fluctuations that exit around
60 $e$-foldings or so before reheating reenter with physical
wavelengths in the range accessible to cosmological observations.
These spectra provide a distinctive signature of inflation. They can
be measured in a variety of different ways including the analysis of
microwave background anisotropies, velocity flows in the universe,
clustering of galaxies and the abundances of gravitationally bound
objects of various types (for reviews, see Efstathiou (1990); Liddle and
Lyth (1993a)).
Until the measurement of large angle microwave background anisotropies
by the COsmic Background Explorer (COBE) satellite (Smoot et al., 1992;
Wright et al., 1992; Bennett et al., 1994, 1996; see White, Scott, and Silk
(1994) for a general discussion of the microwave background), such
observations covered a fairly limited range of scales, and it was
satisfactory to treat the prediction of a generic inflationary scenario as
giving rise to a scale-invariant (Harrison--Zel'dovich) spectrum of density
perturbations (Harrison, 1970; Zel'dovich, 1972) and a negligible amplitude
of gravitational waves (though even then, it was recognized that the
scale-invariance was only approximate (Bardeen et al., 1983)). Since the
detection by COBE, however, the spectra are now constrained over a range of
scales covering some four orders of magnitude from one megaparsec up to
perhaps ten thousand megaparsecs. Moreover, shortly after the COBE
detection, a number of authors reexamined the possibility that a significant
fraction of the signal could be due to gravitational waves (Krauss and
White, 1992; Davis et al., 1992; Salopek, 1992; Liddle and Lyth, 1992;
Lidsey and Coles, 1992; Lucchin, Matarrese, and Mollerach, 1992; Souradeep
and Sahni, 1992; Adams et al., 1993; Dolgov and Silk, 1993).
Thus, the inflationary prediction must now be considered with much
greater care, even in order to deal with {\em present} observations.
At the next level of accuracy, one finds that different inflation
models make different predictions for the spectra, which can be viewed
as differing magnitudes of variation from the scale-invariant result.
In the simplest approximation the spectra are taken to be power-laws.
Hence, modern observations discriminate between different inflationary
models, and are already sufficient to rule out some models completely
(see e.~g.~Liddle and Lyth, 1992) and substantially constrain the
parameter space of others (Liddle and Lyth, 1993a). Future observations
will make even stronger demands on theoretical precision, and will
certainly tightly constrain inflation.
These deviations from highly symmetric situations such as a
scale-invariant spectrum provide an extremely distinctive way of
probing inflation. This is considerably more powerful than employing
historically emphasised predictions such as a spatially flat universe.
Although a spatially flat universe is indeed a typical (but not
inevitable, see e.g., Sasaki et al. (1993); Bucher, Goldhaber, and Turok
(1995)) outcome of inflation, it appears unlikely that this feature
will be unique to inflation. Moreover, the power that observations
such as microwave background anisotropies provides may be sufficient
to override the rather subjective arguments often made against
inflation models because of their apparent `unnaturalness'. Regardless
of whether a model appears natural or otherwise, it should be the
observations which decide whether it is correct or not.
In a wide range of inflationary models, the underlying dynamics is
simply that of a single scalar field --- the {\em inflaton} ---
rolling in some underlying potential. This scenario is generically referred
to as {\em chaotic inflation} (Linde, 1983, 1990b) in reference to its
choice of initial conditions. This picture is widely favored because of its
simplicity and has received by far the most attention to date. Furthermore,
many superficially more complicated models can be rewritten in this
framework. In view of this we shall concentrate on such a type of model
here.
The generation of spectra of density perturbations and gravitational
waves has been extensively investigated in these theories. The usual
strategy is an expansion in the deviation from scale-invariance,
formally expressed as the {\em slow-roll expansion} (Steinhardt and
Turner, 1984; Salopek and Bond, 1990; Liddle, Parsons, and Barrow, 1994). At
the simplest level of approximation, the spectra can be expressed as
power-laws in wavenumber; further accuracy entails calculation of the
deviations from this power-law approximation.
A crucial aspect of the two spectra is that they are not independent.
In a general sense, this is clear since they correspond at the formal
level to two continuous functions that both have an origin in the
single continuous function expressing the scalar field potential. Such
a link was noted in the simplest situation, where the spectra are
approximated by power-laws, by Liddle and Lyth (1992); the general
situation where the two are linked by a {\em consistency equation} was
expounded in Copeland et al. (1993b, henceforth CKLL1), and an
explicit higher-order version of the simplest equation was found by
Copeland et al. (1994a, henceforth CKLL2). If one had complete
expressions for the entire problem, the consistency relation would be
represented as a differential equation relating the two spectra.
However, we shall argue that it is preferable to express the spectra
via an order-by-order expansion. In this case one obtains a finite set
of {\em algebraic} expressions which represent the coefficients of an
expansion of the full differential equation. The familiar situation is
a single consistency equation that relates the gravitational wave
spectral index to the relative amplitudes of the spectra. This is a
result of the lowest-order expansion. The general situation of
multiple consistency equations does not seem to have been expounded
before, though a second consistency equation did appear in Kosowsky and
Turner (1995). In practice, the observational difficulties associated
with measurements of the details of the gravitational wave spectrum
make it extremely unlikely that any but the first consistency equation
shall ever be needed.
Given a particular set of observations of some accuracy, one can
attempt the bold task of reconstructing the inflaton potential from
the observations. In fact, the situation one hopes for is stronger
than a simple reconstruction, the language of which suggests the
possibility of finding a suitable potential regardless of the
observations. With sufficiently good observations, one can first test
whether the consistency equation is satisfied; in situations where
observations make this test non-trivial it provides a very convincing
vindication of the inflationary scenario. Thus emboldened, one could
then go on to use the remaining, nondegenerate, information to
constrain features of the inflaton potential. Figure 1 illustrates
this procedure schematically.
The main obstacle in reconstruction is the limited range of scales
accessible. Although the observations may span up to four orders of
magnitude, the expansion of the universe is usually so fast during
inflation that this typically translates into only a brief range of
scalar field values. One should therefore not overexaggerate the
usefulness of this approach in determining the detailed structure of
physics at high energy, but one should bear in mind that this may be
the only observational information available of any kind at such
energies.
A second obstacle is that one doesn't observe the primordial spectra
directly, but rather after they have evolved considerably. Although this is
a linear problem (except on the shortest scales) and hence computationally
tractable, the evolution necessarily depends on the various cosmological
parameters, such as the expansion rate and the nature of any dark matter.
The form of the initial spectra must be untangled from their influence.
We shall discuss this in some detail in Section \ref{obs}.
Earlier papers discuss two possible ways of treating observational
data. The bolder strategy is to use estimates of the spectra as
functions of scale (Hodges and Blumenthal, 1990; Grishchuk and Solokhin,
1991; CKLL1). In practice, however, this approach founders through the
lack of theoretically derived exact expressions for the spectra
produced by an arbitrary potential. We shall therefore argue in this
review in favor of the alternative approach, which is usually called
perturbative reconstruction (Turner, 1993a; Copeland et al., 1993a;
CKLL1; Turner, 1993b; CKLL2; Liddle and Turner, 1994). In this approach,
the consistency equation and scalar potential are determined as an
expansion about a given point (regarded either as a single scale in
the spectra or as a single point on the potential), allowing
reconstruction of a region of the potential about that point. This has
the considerable advantage that one can terminate the series when
either theoretical or observational knowledge runs out.
The outline of this review is as follows. We devote two Sections to a
review of the inflation driven by a (slowly)
rolling scalar field. We begin by considering the
classical scalar field dynamics and then proceed to discuss the
generation of the spectra of density perturbations and gravitational
waves. Because an accurate derivation of the predicted spectra is
crucial to this programme, we provide a detailed account of the most
accurate calculation presently available, due to Stewart and Lyth
(1993). In Section \ref{first} we consider the simplest possible
scenario allowing reconstruction, and introduce the notion of the
consistency equation. Section \ref{second} reviews the present
state-of-the-art, where next--order corrections are incorporated into
all expressions. One hopes that observational accuracy will justify
this more detailed analysis, though this depends upon which (if any)
inflation model proves correct. Section \ref{genfram} then expands on
this by describing the full perturbative reconstruction framework,
illustrating how much information can be obtained from which
measurements and demonstrating that one can write a hierarchy of
consistency equations. We then briefly illustrate worked examples on
simulated data in Section \ref{obs}. Before concluding, we devote a section
to an examination of other proposals for constraining the inflaton
potential, without using large-scale structure observations.
\section{~~~Inflationary Cosmology and Scalar Fields}
\label{dyn}
\def9.\arabic{equation}{2.\arabic{equation}}
\setcounter{equation}{0}
\subsection{The fundamentals of inflationary cosmology}
Observations indicate that the density distribution in the universe is
nearly smooth on large scales, but contains significant irregularities on
small scales. These correspond to a hierarchy of structures including
galaxies, clusters and superclusters of galaxies.
One of the most important questions that modern cosmology must address is
why the observable universe is almost, but not quite exactly, homogeneous
and isotropic on sufficiently large scales.
The hot big bang model is able to explain the current expansion of the
universe, the primordial abundances of the light elements and
the origin of the cosmic microwave background radiation; for a review of all
these successes see Kolb and Turner (1990). However, this model as it stands
is unable to explain the origin of structure in the universe. This problem
is related to the well known flatness problem (Peebles and Dicke, 1979) and
is essentially a problem of initial data. It arises because the entropy in
the universe is so large, $S \approx 10^{88}$ (Barrow and Matzner, 1977).
One expects this quantity to be of order unity since it is a dimensionless
constant.
This paradox can be made more quantitative in the following way. The
dynamics of a Friedmann--Robertson--Walker (FRW) universe containing
matter with density $\rho$ and pressure $p$ is determined by the Einstein
acceleration equation
\begin{equation}
\label{acc}
\frac{\ddot{a}}{a} = - \frac{4\pi}{3m_{\rm Pl}^2}
(\rho +3p) ,
\end{equation}
the Friedmann equation
\begin{equation}
\label{Fequation}
H^2 = \frac{8\pi}{3m_{\rm Pl}^2} \rho - \frac{k}{a^2},
\end{equation}
and the mass conservation equation
\begin{equation}
\label{mass}
\dot{\rho} +3H(\rho +p) =0 ,
\end{equation}
where $a(t)$ is the scale factor of the universe, $H\equiv \dot{a}/{a}$ is
the Hubble expansion parameter, a dot denotes differentiation with respect
to cosmic time $t$, $m_{\rm Pl}$ is the Planck mass and $k=0,-1, +1$ for
spatially flat, open, or closed cosmologies, respectively. Units are chosen
such that $c=\hbar =1$.
The Friedmann equation (\ref{Fequation}) may be expressed in terms of the
$\Omega$--parameter. This parameter is defined as the ratio of the energy
density of the universe to the critical energy density $\rho_{\rm c}$ that
is just sufficient to halt the current expansion:
\begin{equation}
\Omega \equiv \frac{\rho}{{\rho}_{\rm c}}, \qquad
{\rho}_{\rm c} \equiv \frac{3 m^2_{\rm Pl} H^2}{8\pi}.
\end{equation}
The current observational values for these parameters are ${\rho}_{\rm c} =
1.88h^2 \times 10^{-29} \quad {\rm g} \quad {\rm cm}^{-3}$ and $H_0 =100 h$
km ${\rm s}^{-1}$ ${\rm Mpc}^{-1}$ where conservatively we have
$0.4 \le h \le 0.8$. Eq. (\ref{Fequation}) simplifies to
\begin{equation}
\label{Omegafriedmann}
\Omega -1 =\frac{k}{a^2 H^2},
\end{equation}
and this implies that
\begin{equation}
\label{of}
\frac{\Omega -1}{\Omega} =\frac{3 m^2_{\rm Pl}}{8\pi}
\frac{k}{\rho a^2} .
\end{equation}
Now, for a radiation--dominated universe, the
equation of state is given by $\rho =3p = \pi^2 g_{\rho} T^4/30$
at some temperature $T$, where $g_{\rho} = {\cal{O}} (10^2)$ represents
the total number of relativistic degrees of freedom in the
matter sector at that time. Thus the scale factor grows as $a
(t) \propto t^{1/2}$ when $k=0$ and the expansion rate is given by
\begin{equation}
\label{hub}
H=1.66 g_{\rho}^{1/2} \left( \frac{T^2}{m_{\rm Pl}} \right) =
\frac{1}{2t} .
\end{equation}
Eq. (\ref{hub}) yields the useful expression
\begin{equation}
\label{veryuseful}
\left( \frac{t}{\rm sec} \right) \approx \left( \frac{T}{\rm MeV}
\right)^{-2}
\end{equation}
and substituting Eqs. (\ref{hub}) and (\ref{veryuseful})
into Eq. (\ref{of}) implies that
\begin{equation}
\label{Omega}
\left| \frac{\Omega -1}{\Omega} \right| \approx \frac{10^{43}}{S^{2/3}}
\left( \frac{t}{\rm sec} \right) \approx \frac{10^{37}}{S^{2/3}}
\left( \frac{{\rm GeV}}{T} \right)^2 \,,
\end{equation}
where $S \approx 10^{88}$ is the entropy contained within the present
horizon. The large amount of entropy in the universe therefore implies that
$\Omega$ must have been very close to unity at early times.
Indeed, we find that $\Omega =1 \pm 10^{-16}$ just one second after the
big bang, the time of nucleosynthesis.
The flatness problem is therefore a problem of understanding why the
(classical) initial conditions corresponded to a universe that was so close
to spatial flatness. In a sense, the problem is one of fine--tuning and
although such a balance is possible in principle, one nevertheless feels
that it is unlikely. On the other hand, the flatness problem arises because
the entropy in a comoving volume is conserved. It is possible, therefore,
that the problem could be resolved if the cosmic expansion was
non--adiabatic for some finite time interval $t \in [t_{\rm i},t_{\rm f}]$
during the early history of the universe.
This point was made explicitly by Guth in his seminal paper of 1981. He
postulated that the entropy changed by an amount
\begin{equation}
S_{\rm f}=Z^3S_{\rm i}
\end{equation}
during this time interval, where $Z$ is a numerical factor. In Guth's
original model, this entropy production occurred at, or just below, the
energy scale $T_{\rm GUT} = {\cal{O}} (10^{17})$ GeV associated with the
Grand Unified (GUT) phase transition. This corresponds to a timescale $t
\approx 10^{-40}$ s. Eq. (\ref{Omega}) then implies that the flatness
problem is solved, in the sense that $|{\Omega}_{\rm i}^{-1}-1| = {\cal{O}}
(1)$, if $Z \ge 10^{28}$. It can be shown that the other problems of the big
bang model, such as the horizon and monopole problems are also solved if $Z$
satisfies this lower bound (Guth, 1981).
Guth called this process of entropy production {\em inflation},
because the volume of the universe also grows by the factor $Z^3$
between $t = t_{\rm i}$ and $t = t_{\rm f}$. Indeed, the expansion of the
universe during the inflationary epoch is very rapid. Further insight into
the nature of this expansion may be gained by considering Eq. (\ref{of}).
This expression implies that the quantity $({\Omega}^{-1}-1){\rho}a^2$ is
conserved for an arbitrary equation of state. It follows, therefore, that
\begin{equation}
({\Omega}^{-1}_{\rm i}-1)a_{\rm i}^2{\rho}_{\rm i}=({\Omega}^{-1}_{\rm f}-1)
a_{\rm f}^2{\rho}_{\rm f}
\end{equation}
and, if we assume that the standard, big bang
model is valid for $t>t_{\rm f}$, we may deduce that (Lucchin and
Matarrese, 1985b)
\begin{equation}
{\rho}_{\rm i}a_{\rm i}^2|{\Omega}_{\rm i}^{-1}-1| \approx 10^{-56}
{\rho}_{\rm f}a_{\rm f}^2 |{\Omega}_0^{-1}-1 | .
\end{equation}
Since our current observations imply that $|{\Omega}_0^{-1}-1| = {\cal{O}}
(1)$, the flatness problem is solved if ${\rho}_{\rm f}a_{\rm f}^2 \gg
{\rho}_{\rm i}a_{\rm i}^2$. However, Eq. (\ref{Fequation}) implies that
the quantity $3\dot{a}^2-(8\pi /m^2_{\rm Pl}) \rho a^2$ is also conserved.
Consequently, this inequality is satisfied if $\dot{a}_{\rm f}
>{\dot{a}_{\rm i}}$. Thus, a necessary condition for inflation to proceed is
that the scale factor of the universe {\it accelerates} with respect to
cosmic time:
\begin{equation}
\label{acceleration}
{\ddot a}(t) >0 \,.
\end{equation}
This is in contrast to the decelerating expansion that arises in the big
bang model.
The question now arises as to the nature of the energy source that drives
this accelerated expansion. It follows from Eq. (\ref{acc}) that Eq.
(\ref{acceleration}) is satisfied if $\rho +3p <0$ and this is equivalent to
violating the strong energy condition (Hawking and Ellis, 1973). The
simplest way to achieve such an antigravitational effect is by the presence
of a homogeneous scalar field, $\phi$, with some self--interaction
potential $V(\phi ) \ge 0$. In the FRW universe, such a field is equivalent
to a perfect fluid with energy density and pressure given by
\begin{equation}
\rho =\frac{1}{2} \dot{\phi}^2 + V(\phi)
\end{equation}
and
\begin{equation}
p=\frac{1}{2} \dot{\phi}^2 - V(\phi ) \,,
\end{equation}
respectively. Other matter fields play a negligible role in the evolution
during the inflation, so their presence will be ignored. In this case,
Eqs. (\ref{Fequation}) and (\ref{mass}) are given by
\begin{equation}
\label{F}
H^2 = \frac{8\pi}{3m_{\rm Pl}^2} \left( \frac{1}{2} \dot{\phi}^2
+V(\phi) \right) -\frac{k}{a^2}
\end{equation}
and
\begin{equation}
\label{phieqn}
\ddot{\phi} +3H\dot{\phi} =-V' (\phi) ,
\end{equation}
where here
and throughout a prime denotes differentiation with respect to $\phi$.
Hence, $-\rho \le p \le \rho$ and we have the inflationary requirement
$\ddot{a}>0$ as long as $\dot{\phi}^2 <V$. Inflation is thus achieved
when the matter sector of the theory applicable at some stage in the early
universe is dominated by vacuum energy.
Recently, an alternative inflationary scenario --- the pre--big bang
cosmology --- has been developed whereby the accelerated expansion is
driven by the kinetic energy of a scalar field rather than its
potential energy (Gasperini and Veneziano, 1993a,b, 1994). If the field is
non--minimally coupled to gravity in an appropriate fashion, this kinetic
energy can produce a sufficiently negative pressure and a violation of the
strong energy condition (Pollock and Sahdev, 1989; Levin, 1995a). Such
couplings arise naturally within the context of the string effective
action. However, models of this sort inherently suffer from a `graceful
exit' problem due to the existence of singularities in both the curvature
and the scalar field motion (Brustein and Veneziano, 1994; Kaloper, Madden
and Olive, 1995, 1996; Levin, 1995b; Easther, Maeda, and Wands, 1996).
Moreover, a satisfactory mechanism for generating structure formation and
microwave background anisotropies in these models has yet to be developed,
although it is possible that such inhomogeneities may be generated by
quantum fluctuations in the electromagnetic field (Gasperini, Giovannini and
Veneziano, 1995).
In view of this, we shall restrict our discussion to potential-driven
models. We will focus in this work on some of the general features
of the chaotic inflation scenario (Linde, 1983, 1990b). Although Linde's
original paper considered a specific potential (a quartic one), the theme
was much more general. We adopt the modern usage of {\em chaotic inflation}
to refer to any model where inflation is driven by a single scalar field
slow-rolling from a regime of extremely high potential energy. The phrase
does not imply any particular choice of potential. Most, though not quite
all, modern inflationary models fall under the umbrella of this definition.
Since the precise identity of the scalar field driving the inflation is
unknown, it is usually referred to as the {\em inflaton field}.
In the chaotic inflation scenario, it is assumed that the universe emerged
from a quantum gravitational state with an energy density comparable to that
of the Planck density. This implies that $V (\phi ) \approx m^4_{\rm Pl}$
and results in a large friction term in the Friedmann equation (\ref{F}).
Consequently, the inflaton will slowly roll down its potential, i.e., $|
\ddot{\phi} |\ll H |\dot{\phi} |$ and $\dot{\phi}^2 \ll V$. The condition
for inflation is therefore satisfied and the scale factor grows as
\begin{equation}
a(t) =a_{\rm i} \exp \left( \int^t_{t_{\rm i}} dt' H(t') \right) .
\end{equation}
The expansion is quasi--exponential in nature, since $H(\phi) \approx 8\pi
V(\phi) / 3m_{\rm Pl}^2$ is almost constant, and the curvature term $k/a^2$
in Eq. (\ref{F}) is therefore rapidly redshifted away. The kinetic energy of
the inflaton gradually increases as it rolls down the potential towards the
global minimum. Eventually, its kinetic energy dominates over the potential
energy and inflation comes to an end when $\dot{\phi}^2 \approx V(\phi )$.
The field then oscillates rapidly about the minimum and the couplings of
$\phi$ to other matter fields then become important. It is these
oscillations that result in particle production and a reheating of the
universe.
The simplest chaotic inflation model is that of a free field with a
quadratic potential, $V(\phi) =m^2 \phi^2/2$, where $m$ represents the mass
of the inflaton. During inflation the scale factor grows as
\begin{equation}
a(t) = a_{\rm i} e^{2\pi (\phi^2_{\rm i} - \phi^2 (t))}
\end{equation}
and inflation ends when $\phi = {\cal{O}} (1)$ $ m_{\rm Pl}$. If inflation
begins when $V(\phi_{\rm i} ) \approx m_{\rm Pl}^4$, the scale factor grows
by a factor $\exp( 4\pi m_{\rm Pl}^2/m^2)$ before the inflaton reaches the
minimum of its potential (Linde, 1990b). One can further show that the mass
of the field should be $m \approx 10^{-6}m_{\rm Pl}$ if the microwave
background constraints are to be satisfied. This implies that the volume of
the universe will increase by a factor of $Z^3 \approx 10^{3 \times
10^{12}}$ and this is more than enough inflation to solve the problems of
the hot big bang model.
It is important to emphasize that in this scenario the initial value of the
scalar field is randomly distributed in different regions of the universe.
On the other hand, one need only assume that a small, causally connected,
region of the pre--inflationary universe becomes dominated by the potential
energy of the inflaton field. Indeed, if the original domain is only one
Planck length in extent, its final size will be of the order $10^{10^{12}}$
cm; for comparison, the size of the observable universe is approximately
$10^{28}$ cm.
In conclusion, therefore, the chaotic inflationary scenario represents a
powerful framework within which specific inflationary models can be
discussed. The essential features of each model --- such as the final reheat
temperature and the amplitude of scalar and tensor fluctuations ---
are determined by the specific form of the potential function $V(\phi)$.
This in turn is determined by the particle physics sector of the theory.
Unfortunately, however, there is currently much theoretical
uncertainty in the correct form of the unified field theory above
the electroweak scale. This has resulted in the development of a
large number of different inflationary scenarios and the identity of
the inflaton field is therefore somewhat uncertain. Possible
candidates include the Higgs bosons of grand unified theories, the
extra degrees of freedom associated with higher metric derivatives in
extensions to general relativity, the dilaton field of string theory
and, more generally, the time--varying gravitational coupling that
arises in scalar--tensor theories of gravity.
It is not the purpose of this review to discuss the relative merits of
different models, since this has been done elsewhere (Kolb and Turner, 1990;
Linde, 1990b; Olive, 1990; Liddle and Lyth, 1993a). Traditionally, a
specific potential with a given set of coupling constants is chosen. The
theoretical predictions of the model are then compared with large--scale
structure observations. The region of parameter space consistent with such
observations may then be identified (Liddle and Lyth, 1993a). However, it is
difficult to select a unique inflationary model by this procedure due to the
large number of plausible models available.
In view of the above uncertainties and motivated by recent and forthcoming
advances in observational cosmology, our aim will be to address the question
of whether {\em direct} insight into the nature of the inflaton potential
may be gained by studying the large--scale structure of the universe. We
therefore assume nothing about the potential except that it leads to
an epoch of inflationary expansion.
We will proceed in the remainder of this Section by reviewing a
formalism that allows the classical dynamics of the scalar field
during inflation to be studied in full generality. This formalism may
then be employed to discuss the generation of quantum fluctuations in
the inflaton and gravitational fields.
\subsection{Scalar field dynamics in inflationary cosmology}
In view of the discussion in the previous Subsection, we will assume
throughout this work that the universe was dominated during inflation by a
single scalar field $\phi$ with a self-interaction potential $V(\phi)$, the
form of which it is our aim to determine. We shall further assume that
gravity is adequately described by Einstein's theory of general
relativity. We shall therefore employ the four-dimensional action
\begin{equation}
\label{action}
S= -\int d^4 x \sqrt{-g} \left[ \frac{m_{{\rm Pl}}^2R}{16\pi}
-\frac{1}{2} \left( \nabla {\phi} \right)^2 +V( {\phi})
\right] \,,
\end{equation}
where $R$ is the Ricci curvature scalar of the space--time with metric
$g_{\mu\nu}$ and $g\equiv {\rm det} g_{\mu\nu}$.
Actually, these restrictions are not as strong as they seem. For
example, even theories such as hybrid inflation, which feature
multiple scalar fields, are usually dynamically dominated by only one
degree of freedom (Linde, 1990a, 1991, 1994; Copeland et al., 1994b;
Mollerach, Matarrese, and Lucchin, 1994). Many other models invoke
extensions to general relativity, and much effort has been devoted to
studying inflation in the Bergmann-Wagoner class of generalized
scalar-tensor theories (Bergmann, 1968; Wagoner, 1970) and higher-order
pure gravity theories in which the Einstein-Hilbert lagrangian is
replaced with some analytic function $f(R)$ of the Ricci curvature.
Such theories can normally be rewritten via a conformal transformation
as general relativity plus one or more scalar fields, again with the
possibility that only one such field is dynamically relevant (Higgs,
1959; Whitt, 1984; Barrow and Cotsakis, 1988; Maeda, 1989; Kalara, Kaloper,
and Olive, 1990; Lidsey, 1992; Wands, 1994).
We are unable to discuss models where more than one field is
dynamically important in the reconstruction context. While
considerable progress has been made recently in understanding the
perturbation spectra from these models (Starobinsky and Yokoyama, 1995;
Garc\'{\i}a-Bellido and Wands, 1995, 1996; Sasaki and Stewart, 1996;
Nakamura and Stewart, 1996), the extra freedom of the second field thwarts
any attempt at finding a unique reconstruction, though it is possible to
find some general inequalities relating the spectra (Sasaki and Stewart,
1996). These problems arise both because there is no longer a unique
trajectory, independent of initial conditions, into the minimum of the
potential, and because with a second field one can generate
isocurvature perturbations as well as adiabatic ones. Fortunately, it
appears that it is hard, though not impossible, to keep models of this kind
consistent with observation, as the density perturbations tend to be large
whatever the energy scale of inflation (Garc\'{\i}a-Bellido, Linde, and
Wands, 1996). A completely different way of using two fields is to drive
successive periods of inflation, as in the double inflation scenario
(Polarski and Starobinsky, 1995 and refs therein). This can impose very
sharp features in the spectra which, although rather distinctive, are
not amenable to the perturbative approach that reconstruction
requires.
As we saw above, the accelerated expansion during inflation causes the
spatial hypersurfaces to rapidly tend towards flatness. Moreover, any
initial anisotropies and inhomogeneities in the universe are washed away
beyond currently observable scales by the rapid expansion. Since only the
final stages of the accelerated expansion are important from an
observational point of view, we can assume that the space-time metric may be
described as a spatially flat FRW metric, given by
\begin{equation}
\label{background}
ds^2=L^2(t) dt^2 -e^{2\alpha (t)} [dx^2+dy^2+dz^2] \,,
\end{equation}
where $L(t)$ represents the lapse function and $a(t) =e^{\alpha (t)}$
is the scale factor of the universe.
By taking this metric, we prevent ourselves from studying reconstruction in
the recently discovered versions of inflation giving an open universe (Gott,
1982; Gott and Statler, 1984; Sasaki et al., 1993; Bucher et al., 1995;
Linde, 1995; Linde and Mezhlumian, 1995). In fact these models have not yet
been developed sufficiently to provide the information we need --- in
particular the gravitational wave spectrum has not been predicted --- and
the generalization of the reconstruction program to these models must await
further developments.
Our analysis will however apply in full to low-density cosmological
models where the spatial geometry is kept flat by the introduction of
a cosmological constant (or similar mechanism). Our discussion is entirely
focussed on the initial spectra, which are independent of the material
composition of the universe at late times. Of course, in such a
cosmology the details of going between these spectra and actual
observables will be changed, and the impact of this on reconstruction
has been studied by Turner and White (1995).
Substitution of the metric ansatz Eq.~(\ref{background}) into the
theory given by Eq.~(\ref{action}) leads to an Arnowitt, Deser and
Misner (ADM) (1962) action of the form
\begin{equation}
\label{ADM}
S=\int dt \, U L e^{3\alpha} \left[ -\frac{3m_{{\rm Pl}}^2}{8\pi}
\frac{\dot{\alpha}^2}{L^2} + \frac{1}{2} \frac{\dot{\phi}^2}{L^2}
-V(\phi) \right] \,,
\end{equation}
where $U\equiv \int d^3 {\bf x}$ is the comoving volume of the
universe and a dot denotes differentiation with respect to $t$.
Without loss of generality we may normalize the comoving volume to
unity.
In recent years, considerable progress in the treatment of scalar
fields within the environment of the very early universe has been
made. The approach we adopt in this work is to view the scalar field
itself as the dynamical variable of the system (Grishchuk and Sidorav,
1988; Muslimov, 1990; Salopek and Bond, 1990, 1991; Lidsey, 1991b). This
allows the Einstein-scalar field equations to be written as a set of
first-order, non-linear differential equations.
The Hamiltonian constraint ${\cal{H}}=0$ is derived by functionally
differentiating the action Eq.~(\ref{ADM}) with respect to the
non-dynamical lapse function. One arrives at the Hamilton-Jacobi
equation
\begin{equation}
\label{HJ}
-\frac{4\pi}{3m_{{\rm Pl}}^2} \left( \frac{\partial S}{\partial \alpha}
\right)^2 + \left( \frac{\partial S}{\partial \phi} \right)^2 +2
e^{6\alpha} V(\phi ) =0 \,,
\end{equation}
where the momenta conjugate to $\alpha$ and $\phi$ are $p_{\alpha} =
\partial S/\partial \alpha = -3m_{{\rm Pl}}^2 e^{3\alpha}
\dot{\alpha}/4\pi L$ and $p_{\phi}= \partial S/\partial \phi = e^{3\alpha}
\dot{\phi}/L$, respectively. This equation follows from the invariance
of the theory under reparametrizations of time. The classical dynamics
of this model is determined by the real, separable solution
\begin{equation}
\label{sepact}
S=-\frac{m_{{\rm Pl}}^2}{4\pi} e^{3\alpha} H(\phi) \,,
\end{equation}
where $H(\phi)$ satisfies the differential equation (Grishchuk and
Sidorav, 1988; Muslimov, 1990; Salopek and Bond, 1990, 1991)
\begin{equation}
\label{H}
\left( \frac{dH}{d\phi} \right)^2 -\frac{12\pi}{m_{{\rm Pl}}^2} H^2 (\phi)
=-\frac{32\pi^2}{m_{{\rm Pl}}^4} V(\phi) \,.
\end{equation}
In the gauge $L=1$, substitution of ansatz Eq.~(\ref{sepact}) into the
expressions for the conjugate momenta implies that
\begin{equation}
\label{field}
H(\phi)=\dot{\alpha} \quad ; \quad -\frac{m_{{\rm Pl}}^2}{4\pi}
\frac{dH}{d\phi} = \dot{\phi} \,.
\end{equation}
Thus, $H(\phi)$ represents the Hubble expansion parameter expressed as
a function of the scalar field $\phi$. It follows immediately from the
second of these expressions that $\dot{H}<0$. Consequently, the
physical Hubble radius $H^{-1}$ increases with time as the inflaton
field rolls down its potential. The Hubble radius can only remain
constant if the inflaton field is trapped in a meta-stable false
vacuum state; this is forbidden in the context of `old' inflation as
it can never successfully escape this state, but may be possible in
the context of single-bubble open inflationary models which are
outside the scope of this paper (see, for example, Sasaki et al.,
1993; Bucher et al., 1994; Bucher, Goldhaber, and Turok, 1995; Linde,
1995).
The solution to Eq.~(\ref{H}) depends on an initial condition, the
value of $H$ at some initial $\phi$ (Salopek and Bond, 1990, 1991). If
we are to obtain unique results, the late-time evolution (that is, the
evolution during which the perturbations we see are generated) of $H$
in terms of the scalar field must be independent of the initial
condition chosen, and fortunately one can easily show that this is the
case (Salopek and Bond, 1990; Liddle et al., 1994); the late-time
behavior is governed by an inflationary `attractor' solution, which
is approached exponentially quickly during inflation.
The Hamilton--Jacobi formalism we have outlined is equivalent to the
more familiar version of the equations of motion given by Eqs. (\ref{F}) and
(\ref{phieqn}) (for $k=0$). Eq. (\ref{H}) is equivalent to the time--time
component of the Einstein field equations and therefore represents the
Friedmann equation (\ref{F}). In the form given by Eqs. (\ref{F}) and
(\ref{phieqn}), $\dot{\phi}$ is an initial condition at some value of $t$;
in the Hamilton-Jacobi formalism the equivalent freedom allows one to
specify $H$ at some initial value of $\phi$.
The above analysis of the Hamilton--Jacobi formalism assumes
implicitly that the value of the scalar field is a monotonically
varying function of cosmic time. In particular, it breaks down if the
field undergoes oscillations (though one can attempt to patch together
separate solutions). As a result, this formalism is not directly
suitable for investigating the dynamics of a field undergoing
oscillations in a minimum of the potential, for example. However, the
scalar and tensor fluctuations relevant to large-scale structure
observations are generated when the field is still some distance away
from the potential minimum. Moreover, the piece of the potential
corresponding to these scales is relatively small, so it is reasonable
to assume that the potential is a smoothly decreasing function in this
regime. The scalar field will therefore roll down this part of the
potential in an unambiguous fashion. In the following, we will assume,
without loss of generality, that $\dot{\phi} > 0$, so that $H' (\phi) <0$.
This choice allows us to fix the sign of any prefactors that arise when
square roots appear.
In principle, the Hamilton--Jacobi formalism enables us to treat the
dynamical evolution of the scalar field exactly, at least at the
classical level. In practice, however, the separated Hamilton-Jacobi
equation, Eq.~(\ref{H}), is rather difficult to solve. On the other
hand, the analysis can proceed straightforwardly once the functional
form of the expansion parameter $H(\phi)$ has been determined. This
suggests that one should view $H(\phi)$ as the fundamental quantity in
the analysis (Lidsey, 1991a, 1993). This is in contrast to the more
traditional approaches to inflationary cosmology, whereby the particle
physics sector of the model --- as defined by the specific form of the
inflaton potential $V(\phi)$ --- is regarded as the input parameter.
In the reconstruction procedure, however, the aim is to determine this
quantity from observations, so one is free to choose other quantities
instead. It proves convenient to express the scalar and tensor
perturbation spectra in terms of $H(\phi)$ and its derivatives.
Unfortunately, exact expressions for these perturbations have not yet
been derived in full generality. All calculations to date have
employed some variation of the so-called `slow-roll' approximation
(Steinhardt and Turner, 1984; Salopek and Bond, 1990; Liddle and Lyth, 1992;
Liddle et al., 1994). It is important to emphasize that there are two
different versions of the slow-roll approximation, with their attendant
slow-roll parameters $\epsilon$, $\eta$, etc, depending on whether one is
taking the potential or the Hubble parameter as the fundamental quantity ---
the differences are described in considerable detail in Liddle et al.
(1994). Here we are defining them in terms of the Hubble parameter.
We represent the slow-roll approximation as an expansion in terms of
quantities derived from appropriate derivatives of the Hubble
expansion parameter. Since {\em at a given point} each derivative is
independent, there are in general an infinite number of these terms.
Typically, however, only the first few enter into any expressions
of interest. We define the first three as\footnote{Note that the definition
of the third parameter is different to that made in CKLL2, $\xi_{\rm CKLL2}
= (m_{{\rm Pl}}^2/4\pi) H'''/H'$. The two are related by $\xi^2
=\epsilon \xi_{\rm CKLL2}$. The former definition has proven awkward;
because of the derivative on the denominator it need not be small in
the scale-invariant limit (though the combination $\sqrt{\epsilon
\xi_{{\rm CKLL2}}}$ must be). We choose to use this better definition,
as introduced by Liddle et al. (1994) who give further details and a
collection of useful formulae.}:
\begin{equation}
\label{epsilon}
\epsilon (\phi) \equiv \frac{3\dot{\phi}^2}{2} \left[ V+\frac{1}{2}
\dot{\phi}^2 \right]^{-1} =\frac{m_{{\rm Pl}}^2}{4\pi} \left(
\frac{H' (\phi) }{H(\phi)} \right)^2 \,,
\end{equation}
\begin{equation}
\label{eta}
\eta (\phi)
\equiv -\frac{\ddot{\phi}}{H\dot{\phi}} = \frac{m_{{\rm Pl}}^2}{4\pi}
\frac{H''(\phi)}{H(\phi)} = \epsilon -\frac{m_{{\rm Pl}} \,
\epsilon'}{\sqrt{16\pi \epsilon}} \,,
\end{equation}
\begin{equation}
\label{xi}
\xi (\phi) \equiv \frac{m_{{\rm Pl}}^2}{4\pi} \left( \frac{H'(\phi) H'''
(\phi)}{H^2(\phi)} \right)^{1/2} = \left( \epsilon \eta -
\left( \frac{m_{{\rm Pl}}^2 \, \epsilon}{4\pi} \right)^{1/2}
\eta' \right)^{1/2} \,.
\end{equation}
One need not be concerned as to the sign of the square root in the
definition of $\xi$; it turns out that only $\xi^2$, and not $\xi$
itself, will appear in our formulae (Liddle et al., 1994). We emphasize
that the choice $\dot{\phi} > 0$ implies that $\sqrt{\epsilon} = -
\sqrt{m_{{\rm Pl}}^2/4\pi} \, H'/H$.
Modulo a constant of proportionality, $\epsilon$ measures the relative
contribution of the field's kinetic energy to its total energy
density. The quantity $\eta$, on the other hand, measures the ratio of
the field's acceleration relative to the friction acting on it due to
the expansion of the universe. The slow-roll approximation applies
when these parameters are small in comparison to unity, i.~e.~~$\{
\epsilon, |\eta |, \xi \} \ll 1$; this corresponds to being able to
neglect the first term in Eq.~(\ref{H}) and its first few derivatives.
Inflation proceeds when the scale factor accelerates, $\ddot{a}>0$,
and this is precisely equivalent to the condition $\epsilon <1$.
Inflation ends once $\epsilon$ exceeds unity. It is interesting that
the conditions leading to a violation of the strong energy condition
are uniquely determined by the magnitude of $\epsilon$ alone. In
principle, inflation can still proceed if $|\eta |$ or $|\xi |$ are
much larger than unity, though normally such values would drive a
rapid variation of $\epsilon$ and bring about a swift end to
inflation.
For specific results, we shall not go beyond these three parameters.
However, in general one can define a full hierarchy of slow-roll parameters
(Liddle et al., 1994):
\begin{eqnarray}
\label{generalsro}
\beta_n & \equiv & \left\{\, \prod_{i=1}^{n}\left[- \frac{d \ln
H^{\left(i\right)}}{d \ln a}\right] \right\}^{\frac{1}{n}}
\,, \nonumber \\
& = & \frac{m^{2}_{{\rm Pl}}}{4\pi} \left(\frac{
\left(H'\right)^{n-1}H^{ \left( n+1\right)}}{
H^{n}}\right)^{\frac{1}{n}} \,,
\end{eqnarray}
where $\beta_1 \equiv \eta$, $\beta_2 \equiv \xi$, etc, and a
superscript $(m)$ indicates the $m$-th derivative with respect to
$\phi$. The $\epsilon$ parameter has to be defined separately, though
it may be referred to as $\beta_0$.
These slow-roll parameters, along with analogues defined in terms of
the potential, can be used as the basis for a slow-roll expansion to
derive arbitrarily accurate solutions given a particular choice of
potential. However, this formalism is not necessary when making
general statements about inflation without demanding a specific
potential.
The amount of inflationary expansion within a given timescale is most
easily para\-met\-rized in terms of the number of $e$-foldings that
occur as the scalar field rolls from a particular value $\phi$ to its
value $\phi_e$ when inflation ends:
\begin{equation}
\label{efolds}
N(\phi, \phi_e ) \equiv \int_t^{t_e} H(t) dt =-\frac{4\pi}{m_{{\rm Pl}}^2}
\int^{\phi_e}_{\phi} d\phi \frac{H(\phi)}{H'(\phi)} \,.
\end{equation}
Thus, with the help of Eq.~(\ref{efolds}), we may relate the value of
the scale factor $a(\phi) =e^{\alpha (\phi)}$ at any given epoch
during inflation directly to the value of the scale factor at the end
of inflation, $a_e$:
\begin{equation}
a(\phi) = a_e \exp [-N(\phi)] \,.
\end{equation}
An extremely useful formula is that which connects the two epochs at
which a given scale equals the Hubble radius, the first during
inflation when the scale crosses outside and the second much nearer
the present when the scale crosses inside again. A comoving scale $k$
crosses outside the Hubble radius at a time which is $N(k)$
$e$-foldings from the end of inflation, where
\begin{equation}
\label{Ncross}
N(k) = 62 - \ln \frac{k}{a_0 H_0} - \ln \frac{10^{16}
{\rm GeV}}{V_k^{1/4}}
+ \ln \frac{V_k^{1/4}}{V_{{\rm end}}^{1/4}} - \frac{1}{3} \ln
\frac{V_{{\rm end}}^{1/4}}{\rho_{{\rm reh}}^{1/4}} \,.
\end{equation}
The subscript `0' indicates present values; the subscript `$k$'
specifies the value when the wave number $k$ crosses the Hubble radius
during inflation ($k=aH$); the subscript `end' specifies the value at
the end of inflation; and $\rho_{{\rm reh}}$ is the energy density of
the universe after reheating to the standard hot big bang evolution.
This calculation assumes that instantaneous transitions occur between
regimes, and that during reheating the universe behaves as if
matter-dominated.
It is fairly standard to make a generic assumption about the number of
$e$-foldings before the end of inflation at which the scale presently
equal to the Hubble radius crossed outside during inflation; most
commonly one sees this number taken as either 50 or 60. Within the
context of making predictions from a given potential this can have a
slight effect on results, but it is completely unimportant as regards
reconstruction.
What we do need for reconstruction is a measure of how rapidly scales
pass outside the Hubble-radius as compared to the evolution of the
scalar field; this is essential for calculating such quantities as the
spectral indices of scalar and tensor perturbations. The formal
definition we take of a scale matching the Hubble radius is that $k =
aH$. Then one can write
\begin{equation}
\label{scalescalar}
k(\phi) = a_e H(\phi) \exp[ -N(\phi) ] \,,
\end{equation}
where $N(\phi)$ is given by Eq.~(\ref{efolds}). Differentiating with
respect to $\phi$ therefore yields
\begin{equation}
\label{kphi}
\frac{d \ln k}{d \phi} =\frac{4\pi}{m_{{\rm Pl}}^2} \frac{H}{H'}
(\epsilon - 1) \,.
\end{equation}
This concludes our discussion on the classical dynamics of the scalar
field during inflation. In the following Section, we will proceed to
discuss the consequences of quantum fluctuations that arise in both
the inflaton and graviton fields.
\section{~~~The Quantum Generation of Perturbations}
\label{pert}
\def9.\arabic{equation}{3.\arabic{equation}}
\setcounter{equation}{0}
During inflation, the inflaton and graviton fields undergo
quantum-mechanical fluctuations. The most important observational
consequences of the inflationary scenario derive from the significant
effects these perturbations may have on the large-scale structure of
the universe at the present epoch. In this Section we shall discuss
how these fluctuations arise and present expressions for their
expected amplitudes. Since the inflaton and gravitational
perturbations are produced in a similar fashion, we shall begin with a
qualitative description of the effects of the former. We shall then
proceed with an extensive account of the calculation of both spectra
by Stewart and Lyth (1993), which is the most accurate analytic
treatment presently available.
\subsection{Qualitative discussion}
Fluctuations in the inflaton field lead to a stochastic spectrum of
density (scalar) perturbations (Guth and Pi, 1982; Hawking, 1982; Linde,
1982b; Starobinsky, 1982; Bardeen et al., 1983; Lyth, 1985; Mukhanov,
1985; Sasaki, 1986; Mukhanov, 1989; Salopek, Bond, and Bardeen, 1989).
Physically, these arise because the inflaton field reaches the global
minimum of its potential at different times in different places in the
universe. This results in a time shift in how quickly the rollover
occurs. Thus, constant $\delta\rho$ does not correspond to a
constant-time hypersurface; in other words, there is a density
distribution produced by the kinetic energy of the inflaton field for
a given constant-time hypersurface. It is widely thought that these
density perturbations result in the formation of large-scale structure
in the universe via the process of gravitational instability. They may
also be responsible for anisotropic structure in the temperature
distribution of the cosmic microwave background radiation.
Typically, the inflationary scenario predicts that the spectrum of
density perturbations should be gaussian and scale-dependent. This is
certainly true for the class of models that we shall be considering
here, in which it is assumed that the inflaton field is weakly
coupled. However, one should bear in mind that the prediction of
gaussianity is not generic to all inflationary models; it is possible
to contrive models with nongaussian perturbations by introducing
features in just the right part of the inflationary potential (Allen,
Grinstein, and Wise, 1987; Salopek and Bond, 1990).
The historical viewpoint on the scale-dependence of the fluctuations
was that they were of scale-invariant (Harrison--Zel'dovich) form, though
it had been recognized that the scale-invariance was only approximate
(Bardeen et al., 1983; Lucchin and Matarrese, 1985a). This is because the
scalar field must be undergoing some kind of evolution if inflation is to
end eventually, and this injects a scale-dependence into the spectra. As we
shall see, this effect should be easy to measure.
To take advantage of accurate observations, it is imperative that the
spectra be calculated as accurately as possible. However, let us first
make a qualitative discussion of the generation mechanism.
In a spatially flat, isotropic and homogeneous universe, the Hubble
radius, $H^{-1}(t)$, represents the scale beyond which causal
processes cannot operate. The relative size of a given scale to this
quantity is of crucial importance for understanding how the primordial
spectrum of fluctuations is generated. Quantities such as the power
spectrum are defined via a Fourier expansion as functions of {\em
comoving} wavenumber $k$, and the combination $k/aH$ appears in many
equations. Different physical behavior occurs depending on whether
this quantity is much greater or smaller than unity.
Inflation is defined as an epoch during which the scale factor
accelerates, and so the comoving Hubble radius, $(aH)^{-1}$, must
necessarily decrease. This is an important feature of the
inflationary scenario, because it means that physical scales will grow
more rapidly than the Hubble radius. As a result, a given mode will
start within the Hubble radius. In this regime the expansion is
negligible and the microphysics in operation at that epoch will
therefore be relevant. This is determined by the usual flat-space
quantum field theory for which the vacuum state of the scalar field
fluctuations is well understood. As the inflationary expansion
proceeds, however, the mode grows much more rapidly than the Hubble
radius (in physical coordinates) and soon passes outside it. One can
utilize a Heisenberg picture of quantum theory to say that the
operators obey the classical equations of motion, and so the evolution
of the vacuum state can be followed until it crosses outside the
Hubble radius. At this point the microphysics effectively becomes
`frozen'. It turns out that the asymptotic state is not a
zero-particle state --- particles are created by the gravitational
field. Corresponding perturbations in the gravitational field itself
are also generated, so a spectrum of gravitational wave (tensor)
fluctuations is independently produced by the same mechanism.
Once inflation is over, the comoving Hubble radius begins to grow.
Eventually, therefore, the mode in question is able to come back
inside the Hubble radius some time after inflation. The overall result
is that perturbations arising from fluctuations in the inflaton field
can be imprinted onto a given length scale during the inflationary
epoch when that scale first leaves the Hubble radius. These will be
preserved whilst the mode is beyond the Hubble radius and will
therefore be present when the scale re-enters during the
radiation-dominated or matter-dominated eras.
\subsection{Quantitative analysis}
If one is to take full advantage of the observations to the extent one
hopes, it is crucial to have extremely accurate predictions for the
spectra induced by different inflationary models. For example,
microwave background theorists have set themselves the stringent goal
of calculating the radiation angular power spectrum (the $C_l$
discussed later in this paper) to within one percent (Hu et al., 1995),
in the hope that satellite observations may one day provide
extremely accurate measurements of the anisotropies across a wide
range of angular scales (Tegmark and Efstathiou, 1996). This involves a
detailed treatment with a host of subtle physical effects. If
inflationary models are to capitalize on this sort of accuracy, it is
essential to have as accurate a determination as possible of the
initial spectra which are to be input into such calculations. Given
that the slow-roll parameters are typically at least a few percent,
that implies that a determination of the spectra to at least one order
beyond leading order in the slow-roll expansion is desired.
The calculations we make are based on linear perturbation theory.
Since the observed anisotropies are small, this approximation is
considerably more accurate than the slow-roll approximation, and we
need not attempt to go beyond it, though it is possible to extend
calculations beyond linear perturbation theory (Durrer and
Sakellariadou, 1994).
Before proceeding, however, let us clarify a notational point. In
earlier literature, especially CKLL2 and Liddle and Turner (1994),
orders were referred to as first-order, second-order etc. However, we
feel this can be misleading, because it might suggest that all terms
containing say two slow-roll parameters in any given expression are
supposed to be neglected. This is not the intention, because in many
expressions the lowest-order term already contains one or more powers
of the slow-roll parameters. Because differentiation respects the
order-by-order expansion, while multiplying each term by a slow-roll
parameter, it is always valid to take terms to the same number of
orders, however many slow-roll parameters the actual terms possess.
Therefore, in order to clarify the meaning, we choose to always employ
the phrase \underline{lowest-order} to indicate the term containing
the least number of powers of the slow-roll parameters, however many
that may be for a specific expression. The phrase next-to-lowest
order, abbreviated to \underline{next-order}, then indicates
correction terms to this which contain one further power of the
slow-roll parameters than the lowest-order terms.
The calculation of the spectra to next-order has been provided by
Stewart and Lyth (1993). Because of its crucial importance, we shall
devote quite some time to describing it. The basic principle is to
start with the one known situation where the spectra can be calculated
exactly, that of power-law inflation. This corresponds to each of the
slow-roll parameters having the same constant value. To next-order, a
general inflationary potential can be considered via an expansion in
$(\epsilon - \eta)$ about a power-law inflation model with the same
$\epsilon$; as we shall see, it is an adequate approximation to treat
$\epsilon$ and $\eta$ as different constant values.
In fact, the logic we develop is slightly different to that of Stewart
and Lyth (1993). They computed an exact solution for the situation
where $\epsilon$ and $\eta$ are treated as exactly constant with
different values. Formally, this situation does not exist as
$\epsilon$ precisely constant implies $\epsilon = \eta$. They then
treated power-law inflation as an exact special case of this
situation, and a general inflation model to next-order as an expansion
about their more general result. Logically, it is more accurate to
expand directly about the exact power-law inflation result, but
nevertheless the final answer is guaranteed to be the same.
\subsubsection{Scalar perturbations}
Throughout the calculations to derive the spectra of scalar and tensor
fluctuations, the space-time representing our universe is decoupled
into two components, representing the background and perturbation
contributions. The background part is taken to be the homogeneous and
isotropic FRW metric. This is a reasonable assumption to make in view
of the high degree of spatial uniformity in the temperature of the
cosmic microwave background. In this paper we assume the background is
also spatially flat with a line element given by
Eq.~(\ref{background}). The perturbed sector of the metric then
determines by how much the actual universe deviates from this
idealization.
Four quantities are required to specify the general nature of a scalar
perturbation. These may be denoted by $A$, $B$, $\Psi$ and $E$ and
these are functions of the space and time coordinates. It has been
shown by Bardeen (1980) and by Kodama and Sasaki (1984) that the most
general form of the line element for the background and scalar metric
perturbations is given by
\begin{equation}
\label{metric}
ds^2 =a^2(\tau) \left[ (1+2A)d\tau^2 -2\partial_i B dx^i d\tau -
\left[ (1-2\Psi ) \delta_{ij} +2 \partial_i \partial_j E \right] dx^i dx^j
\right] \,,
\end{equation}
where $\tau \equiv \int dt/a(t)$ is conformal time.
The perturbations can be measured by the intrinsic curvature perturbation of
the comoving hypersurfaces, which has the form
\begin{equation}
\label{intrinsic}
{\cal{R}} = -\Psi -\frac{H}{\dot{\phi}} \delta \phi\,,
\end{equation}
during inflation, where $\delta\phi$ represents the fluctuation of the
inflaton field and $\dot{\phi}$ and $H$ are calculated from the
background field equations Eqs.~(\ref{H})-(\ref{field}). To proceed,
we follow Mukhanov, Feldman and Brandenberger (1992) and introduce the
gauge-invariant potential
\begin{equation}
\label{v}
u \equiv a\left[ \delta \phi +\frac{\dot{\phi}}{H} \Psi \right] \,.
\end{equation}
It also proves convenient to introduce the variable
\begin{equation}
\label{z}
z\equiv \frac{a\dot{\phi}}{H} \,,
\end{equation}
and it follows immediately that
\begin{equation}
\label{vR}
u=-z{\cal{R}} \,.
\end{equation}
The evolution of the perturbations is determined by the Einstein
action. The first-order perturbation equations of motion are given by
a second-order action. Hence, the gravitational and matter sectors are
separated and each expanded to second-order in the perturbations. The
result for the gravitational component is simplified by employing the
ADM form of the action (Arnowitt, Deser, and Misner, 1962; Misner et al.,
1973). The action for the matter perturbations, on the other hand, can be
calculated by expanding the Lagrangian as a Taylor series about a fixed
value of the scalar field, applying the background field equations and
integrating by parts. Mukhanov et al. (1992) show that the full action for
linear scalar perturbations is given by
\begin{equation}
\label{pertact}
S=\int d^4x{\cal{L}} = \frac{1}{2} \int d\tau d^3{\bf x} \left[
\left( \partial_{\tau} u \right)^2 -\delta^{ij} \partial_i
u\partial_j u + \frac{z_{\tau\tau}}{z} u^2 \right] \,,
\end{equation}
where a subscript $\tau$ denotes partial differentiation with respect
to conformal time. For further details the reader is referred to
Mukhanov (1989) and Makino and Sasaki (1991).
Formally, this is equivalent to the action for a scalar field in flat
space-time with a time-dependent effective mass $m^2 =-
z_{\tau\tau}/z$. This equivalence implies that one can consider the
quantum theory in an analogous fashion to that of a scalar field
propagating on Minkowski space-time in the presence of a time-varying
external field (Grib, Mamaev, and Mostepanenko, 1980). The
time-dependence has its origin in the variation of the background
space-time (Birrell and Davies, 1982).
The momentum canonical to $u$ is given by
\begin{equation}
\label{canmom}
\pi (\tau ,{\bf x}) =\frac{\partial {\cal{L}}}{\partial (u_{\tau})}
=u_{\tau} (\tau , {\bf x}) \,,
\end{equation}
and the theory is then quantized by promoting $u$ and its conjugate
momentum to operators that satisfy the following commutation relations
on the $\tau ={\rm constant}$ hypersurfaces:
\begin{eqnarray}
\label{ETCR}
\left[ \hat{u} (\tau ,{\bf x}) ,\hat{u} (\tau ,{\bf y}) \right] =
\left[ \hat{\pi} (\tau ,{\bf x}) ,\hat{\pi} (\tau ,{\bf y}) \right] =0 \,,
\\
\left[ \hat{u} (\tau ,{\bf x}) ,\hat{\pi} (\tau ,{\bf y})
\right] = i\delta^{(3)} ({\bf x} -{\bf y}) \,.
\end{eqnarray}
We expand the operator $\hat{u} (\tau ,{\bf x})$ in terms of plane waves
\begin{equation}
\label{vhat}
\hat{u} (\tau ,{\bf x}) = \int \frac{d^3{\bf k}}{(2\pi)^{3/2}} \left[
u_k (\tau) \hat{a}_{\bf k} e^{i{\bf k.x}} +u_k^* (\tau)
\hat{a}^{\dagger}_{\bf k} e^{-i{\bf k.x}} \right] \,,
\end{equation}
and the field equation for the coefficients $u_k$ is derived by
setting the variation of the action Eq.~(\ref{pertact}) with respect
to $u$ equal to zero. It is given by (Mukhanov, 1985, 1988; Stewart and
Lyth, 1993)
\begin{equation}
\label{vfield}
\frac{d^2u_k}{d\tau^2} +\left( k^2 - \frac{1}{z} \frac{d^2 z}{d\tau^2}
\right) u_k =0 \,.
\end{equation}
These modes are normalized so that they satisfy the Wronskian condition
\begin{equation}
\label{norm}
u^*_k \frac{du_k}{d\tau} - u_k \frac{du^*_k}{d\tau} =-i \,,
\end{equation}
and this condition ensures that the creation and annihilation
operators $\hat{a}^{\dagger}_{\bf k}$ and $\hat{a}_{\bf k}$ satisfy
the usual commutation relations for bosons:
\begin{equation}
\label{usual}
[\hat{a}_{\bf k}, \hat{a}_{\bf l}]=[\hat{a}^{\dagger}_{\bf k},
\hat{a}^{\dagger}_{\bf l}]=0, \qquad [\hat{a}_{\bf k},
\hat{a}^{\dagger}_{\bf l}]=\delta^{(3)} ({\bf k}-{\bf l}) \,.
\end{equation}
The vacuum is therefore defined as the state that is annihilated by
all the $\hat{a}_{\bf k}$, i.~e., $\hat{a}_{\bf k} | 0 \rangle =0$.
The modes $u_k(\tau)$ must have the correct form at very short
distances so that ordinary flat space-time quantum field theory is
reproduced. Thus, in the limit that $k/aH \rightarrow \infty$, the
modes should approach plane waves of the form
\begin{equation}
\label{short}
u_k(\tau) \rightarrow \frac{1}{\sqrt{2 k}} e^{-ik\tau}\,.
\end{equation}
In the opposite (long wavelength) regime where $k$ can be neglected in
Eq.~(\ref{vfield}), we see immediately that the growing mode solution
is
\begin{equation}
\label{long}
u_k \propto z \,,
\end{equation}
with no dependence on the behavior of the scale factor (except
insofar as implicitly through the definition of $z$).
Ultimately, the quantity in which we are interested is the curvature
perturbation ${\cal R}$. We expand this in a Fourier series
\begin{equation}
\label{fourierR}
{\cal{R}} =\int \frac{d^3{\bf k}}{(2\pi)^{3/2}} {\cal{R}}_{\bf k}
(\tau) e^{i{\bf k.x}} \,.
\end{equation}
The power spectrum ${\cal P}_{\cal{R}} (k)$ can then be defined in terms
of the vacuum expectation value
\begin{equation}
\label{scalarpower}
\langle {\cal{R}}_{\bf k} {\cal{R}}^*_{\bf l} \rangle =
\frac{2\pi^2}{k^3} {\cal P}_{\cal{R}} \delta^{(3)} ({\bf k}-{\bf l}) \,,
\end{equation}
where the prefactor is in a sense arbitrary but is chosen to obey the
usual Fourier conventions. The left-hand side of this expression may
be evaluated by combining Eqs.~(\ref{vR}), (\ref{usual}) and
(\ref{fourierR}):
\begin{equation}
\label{vev}
\langle {\cal{R}}_{\bf k}{\cal{R}}^*_{\bf l} \rangle =\frac{1}{z^2}
|u_k|^2 \delta^{(3)} ({\bf k}-{\bf l}) \,,
\end{equation}
yielding
\begin{equation}
\label{pspec}
{\cal P}_{\cal R}^{1/2}(k) = \sqrt{\frac{k^3}{2\pi^2}} \, \left|
\frac{u_k}{z} \right| \,.
\end{equation}
For modes well outside the horizon, the growing mode of $u_k$ will
dominate and so the spectrum will approach a constant value. It is
this value that we are aiming to calculate.
In order to provide a solution, we need an expression for
$z_{\tau\tau}/z$. This can be straightforwardly obtained as
\begin{equation}
\label{zderiv}
\frac{1}{z} \frac{d^2z}{d\tau^2} =2a^2 H^2 \left[ 1+\epsilon -\frac{3}{2}
\eta + \epsilon^2 -2 \epsilon \eta + \frac{1}{2} \eta^2 + \frac{1}{2} \xi^2
\right] \,,
\end{equation}
and despite its appearance as an expansion in slow-roll parameters,
this expression is exact.
\vspace*{12pt}
\noindent
{\bf Exact solution for power-law inflation}
\vspace*{12pt}
So far, all the expressions we have written down have been exact.
However, we have reached the limit of analytic progress for general
circumstances. The desired situation then is to obtain an exact
solution for some special case, about which a general expansion can be
applied in terms of the slow-roll parameters. Such an exact solution
is the case of power-law inflation, which we now derive\footnote{It is
at this point that our construction of the expansion begins to differ
in logical construction from Stewart and Lyth (1993), though the final
result will agree.}.
Power-law inflation, where the scale factor expands as $a(t) \propto
t^p$, corresponds to the particularly simple case where the Hubble
parameter is exponential in $\phi$ (Lucchin and Matarrese, 1985a,
1985b):
\begin{equation}
H(\phi) \propto \exp \left( \sqrt{\frac{4\pi}{p}} \,
\frac{\phi}{m_{{\rm Pl}}} \right) \,.
\end{equation}
It follows that the slow-roll parameters are not only constant but
equal; we are primarily interested in
\begin{equation}
\epsilon = \eta = \xi = \frac{1}{p} \,.
\end{equation}
With a constant $\epsilon$, an integration by parts
\begin{equation}
\label{parts}
\tau = \int \frac{da}{a^2H} = -\frac{1}{aH} + \int \frac{\epsilon \,
da}{a^2H} \,,
\end{equation}
supplies the conformal time as
\begin{equation}
\label{simply}
\tau =-\frac{1}{aH}\frac{1}{1-\epsilon} \,.
\end{equation}
Thus, $\tau$ is negative during inflation, with $\tau = 0$
corresponding to the infinite future.
Since the slow-roll parameters in Eq.~(\ref{zderiv}) are constant,
Eq.~(\ref{vfield}) simplifies to a Bessel equation of the form
\begin{equation}
\label{bessel}
\left[ \frac{d^2}{d\tau^2} +k^2 -\frac{(\nu^2 -\frac{1}{4})}{\tau^2}
\right] u_k =0 \,,
\end{equation}
where
\begin{equation}
\nu \equiv \frac{3}{2} + \frac{1}{p-1} \,.
\end{equation}
The appropriately normalized solution with the correct asymptotic
behavior at small scales is therefore given by\footnote{The choice of
phase factor ensures that the behavior described by Eq.~(\ref{short})
is reproduced at short scales, and the factor of $\sqrt{\pi}/2$
implies that condition Eq.~(\ref{norm}) is satisfied.}
\begin{equation}
\label{correctform}
u_k(\tau) =\frac{\sqrt{\pi}}{2} e^{i(\nu +1/2)\pi/2} (-\tau)^{1/2}
H_{\nu}^{(1)} (-k\tau) \,,
\end{equation}
where $H_{\nu}^{(1)}$ is the Hankel function of the first kind of order
$\nu$.
Ultimately, we are interested in the asymptotic form of the solution
once the mode is well outside the horizon. Taking the limit $k/aH
\rightarrow 0$ yields the asymptotic form
\begin{equation}
\label{modelimit}
u_k \rightarrow e^{i(\nu -1/2)\pi /2} 2^{\nu -3/2}
\frac{ \Gamma (\nu)}{\Gamma (3/2)} \frac{1}{\sqrt{2k}} (-k \tau )^{-\nu
+1/2} \,,
\end{equation}
and substituting this into Eq.~(\ref{pspec}) gives the asymptotic form
of the power spectrum
\begin{equation}
\label{scalaramp}
{\cal P}_{\cal{R}}^{1/2}(k) =2^{\nu -1/2} \frac{\Gamma(\nu)}{\Gamma(3/2)}
(\nu-1/2)^{1/2 - \nu} \frac{1}{m_{{\rm Pl}}^2} \left.
\frac{H^2}{|H'|} \right|_{k=aH} \,,
\end{equation}
where we have employed Eq.~(\ref{simply}) to substitute for $k\tau$.
A subtle point is that, despite the appearance of this equation, the
calculated value for the spectrum is {\em not} the value when the
scale crosses outside the Hubble radius. Rather, it is the asymptotic
value as $k/aH \rightarrow 0$, but rewritten in terms of the values
which quantities had at Hubble radius crossing.
This exact expression for the asymptotic power spectrum was first
derived in an earlier paper by Lyth and Stewart (1992). It is one of
only two known exact solutions, and is the only one for a realistic
inflationary scenario. The other known exact solution, found by
Easther (1996), arises in an artificial model designed to permit exact
solution, and while of theoretical interest is excluded by observations.
\vspace*{12pt}
\noindent
{\bf Slow-roll expansion for general potentials}
\vspace*{12pt}
Having obtained an exact solution, we can now make an expansion about
it. The power-law inflation case corresponded to the slow-roll
parameters being equal, and hence exactly constant; we now wish to
allow them to be different which means they will pick up a time
dependence.
At this stage, there is no need to require that the parameter
$\epsilon$ be small, for the exact solution exists for all $\epsilon <
1$. However, the deviation of all higher slow-roll parameters from
$\epsilon$ must indeed be small, since the differences vanish for the
exact solution. Let us label the first of these as $\zeta = \epsilon -
\eta$. There are in general an infinite number of such small
parameters in the expansion but we shall only need this one.
The first step is to find a more general equation for $\tau$. By
integrating by parts in the manner of Eq.~(\ref{parts}) an infinite
number of times, one can obtain
\begin{equation}
\tau =-\frac{1}{aH}\frac{1}{1-\epsilon} - \frac{2\epsilon \zeta}{aH} +
\mbox{expansion in slow-roll parameters $\zeta$ etc.} \,,
\end{equation}
where $\epsilon$ can now have arbitrary time dependence. This is all
very well, but even via an expansion in small $\zeta$ one cannot
analytically solve Eq.~(\ref{vfield}) for a general time-dependent
$\epsilon$; we must resort to a situation where $aH\tau$ can be taken
as constant for each $k$-mode (though not necessarily the same
constant for different $k$). The relevant equation to study is the
exact relation
\begin{equation}
\label{vareps}
\dot{\epsilon}/H = 2 \epsilon \zeta \,.
\end{equation}
What we are aiming to do is to shift the time dependence of $\epsilon$
to next-order in the expansion, so that it can be neglected. This is
achieved by assuming that $\epsilon$ is a small parameter as well as
$\zeta$ (that is, that both $\epsilon$ and $\eta$ are small), in which
case one can expand to lowest-order to get
\begin{equation}
\tau = -\frac{1}{aH} (1+\epsilon) \,.
\end{equation}
We will return to the question of the error in assuming constant
$\epsilon$ shortly.
Having this expression for $\tau$, we can now immediately use
Eq.~(\ref{zderiv}), which must also be truncated to first-order. This
gives the same Bessel equation Eq.~(\ref{bessel}), but now with $\nu$
given by
\begin{equation}
\nu = \frac{3}{2} + 2\epsilon - \eta \,.
\end{equation}
The assumption that treats $\epsilon$ as constant also allows $\eta$
to be taken as constant, but crucially, $\epsilon$ and $\eta$ need no
longer be the same since we are consistent to first-order in their
difference. The differences between further slow-roll parameters and
$\epsilon$ lead to higher order effects, and so incorporating
$\epsilon$ and $\eta$ in this manner is applicable to an arbitrary
inflaton potential to next-order. The same solution
Eq.~(\ref{scalaramp}) can be used with the new form of $\nu$, but for
consistency it should be expanded to the same order. This gives the
final answer, which is true for general inflation potentials to this
order, of (Stewart and Lyth, 1993)
\begin{equation}
\label{2ndscal}
P_{\cal{R}}^{1/2}(k) = \left[ 1 - (2C+1) \epsilon + C \eta
\right] \; \frac{2}{m_{{\rm Pl}}^2} \left.
\frac{H^2}{|H'|} \right|_{k=aH} \,,
\end{equation}
where $C = -2 +\ln 2 +\gamma \simeq -0.73$ is a numerical constant,
$\gamma$ being the Euler constant originating in the expansion of the
Gamma function. Since the slow-roll parameters are to be treated as
constant, they can also be evaluated at horizon crossing.
Let us now return to the question of the error in assuming $\epsilon$
is constant. The crucial aspect is that the variation of $\epsilon$ is
only important around $k = aH$. In either of the two extreme regimes
the evolution of $u_k$ (in relation to $z$) is independent of it
(Eqs.~(\ref{short}) and (\ref{long})). Assuming the variation of
$\epsilon$ is only important for some unspecified but finite number of
$e$-foldings, Eq.~(\ref{vareps}) measures that change (per
$e$-folding). As long as we are assuming $\epsilon$ small as well as
$\zeta$, that change is next-order and can be neglected along with all
the other next-order terms we did not attempt to include.
Finally, one can see from the complexity of this calculation the
obstacles to obtaining general expressions which go to yet another
higher order. This would involve finding some way of solving the
Bessel-like equation in the situation where its coefficients could not
be treated as constant.
This concludes our discussion on the generation of scalar
perturbations during inflation. In the remainder of this Section we
will present the analogous result for the tensor fluctuations.
\subsubsection{Gravitational waves}
The propagation of weak gravitational waves on the FRW background was
investigated by Lifshitz (1946). Quantum fluctuations in the
gravitational field are generated in a similar fashion to that of the
scalar perturbations discussed above. A gravitational wave may be
viewed as a ripple in the background space-time metric
Eq.~(\ref{background}) and in general the linear tensor perturbations
may be written as $g_{\mu\nu} =a^2 (\tau)[ \eta_{\mu\nu}
+h_{\mu\nu}]$, where $|h_{\mu\nu}| \ll 1$ denotes the metric
perturbation and $\eta_{\mu\nu}$ is the flat space-time metric
(Bardeen, 1980; Kodama and Sasaki, 1984). In the transverse-traceless
gauge, we have $h_{00} = h_{0i} = \partial^i h_{ij} =\delta^{ij}
h_{ij} =0$, and there are two independent polarization states (Misner et
al., 1973). These are usually denoted as $\lambda = +,\times$.
The gravitons are the propagating modes associated with these two
states. The classical dynamics of the gravitational waves is
determined by expanding the Einstein-Hilbert action to quadratic order
in $h_{\mu\nu}$ and it can be shown that this action takes the form
(Grishchuk, 1974, 1977)
\begin{equation}
\label{gravitonaction}
S_g= \frac{m_{{\rm Pl}}^2}{64\pi} \int d\tau d^3 {\bf x} a^2 (\tau)
\partial_{\mu} {h^i}_j \partial^{\mu}{h_i}^j \,.
\end{equation}
It proves convenient to introduce the rescaled variable
\begin{equation}
{P^i}_j (x) \equiv (m_{{\rm Pl}}^2 /32 \pi)^{1/2} a(\tau) {h^i}_j (x) \,,
\end{equation}
and substitution of this expression into the action
Eq.~(\ref{gravitonaction}) implies that
\begin{equation}
\label{newact}
S_g =\frac{1}{2} \int d\tau d^3{\bf x} \left[ \left( \partial_{\tau}
{P_i}^j \right) \left( \partial^{\tau} {P^i}_j \right)
- \delta^{rs} \left( \partial_r {P_i}^j \right) \left( \partial_s
{P^i}_j \right) + \frac{a_{\tau\tau}}{a} {P_i}^j {P^i}_j
\right] \,,
\end{equation}
where we have ignored a total derivative. This expression resembles
the equivalent action Eq.~(\ref{pertact}) for the scalar
perturbations. Indeed, we may interpret Eq.~(\ref{newact}) as the
action for two scalar fields in Minkowski space-time each with an
effective mass squared given by $a_{\tau\tau}/a $. This equivalence
between the two actions implies that the procedure for quantizing the
tensor fluctuations is essentially the same as in the scalar case.
We perform a Fourier decomposition of the gravitational waves by
expanding ${P^i}_j$:
\begin{equation}
{P^i}_j = \sum_{\lambda =+,\times} \int \frac{d^3 {\bf k}}{(2\pi )^{3/2}}
v_{{\bf k}, \lambda} (\tau) {\epsilon^i}_j ({\bf k};\lambda )
e^{i {\bf k.x}} \,.
\end{equation}
In this expression ${\epsilon^i}_j ({\bf k}; \lambda )$ is the
polarization tensor and satisfies the conditions $\epsilon_{ij}
=\epsilon_{ji}$, $\epsilon_{ii}=0$, $k^i\epsilon_{ij}=0$ and
${\epsilon^i}_j ({\bf k},\lambda ) {\epsilon_i}^{j*} ({\bf k},
\lambda' ) =\delta_{\lambda\lambda'}$. The analysis is further
simplified if we choose $\epsilon_{ij}(-{\bf k}, \lambda
)=\epsilon_{ij}^* ({\bf k} ,\lambda )$, since this ensures that
$v_{{\bf k} ,\lambda} =v^*_{-{\bf k} ,\lambda}$. We may consider each
polarization state separately. The effective graviton action during
inflation therefore takes the form
\begin{equation}
\label{rewrite}
S_g=\frac{1}{2} \sum_{\lambda =+,\times} \int d\tau d^3{\bf k}
\left[ \left( \partial_{\tau} \left| v_{{\bf k},\lambda} \right| \right)^2
-\left( k^2 -\frac{a_{\tau\tau}}{a} \right) \left| v_{{\bf k},\lambda}
\right|^2 \right] \,.
\end{equation}
We quantize by interpreting $v_{{\bf k},\lambda} (\tau)$ as the operator
\begin{equation}
\label{quantize}
\hat{v}_{{\bf k},\lambda} (\eta) =v_k(\eta) \hat{a}_{{\bf k},\lambda}
+v^*_k(\eta) \hat{a}^{\dagger}_{-{\bf k},\lambda} \,,
\end{equation}
where the modes $v_k$ satisfy the normalization condition
Eq.~(\ref{norm}) and have the form given by Eq.~(\ref{short}) as $aH/k
\rightarrow 0$. This ensures that the creation and annihilation
operators satisfy
\begin{equation}
\label{gravityETCR}
[\hat{a}_{{\bf k},\lambda} \hat{a}^{\dagger}_{{\bf l},\sigma} ]
=\delta_{\lambda\sigma} \delta^{(3)} ({\bf k}-{\bf l}),
\qquad \hat{a}_{{\bf k},\lambda} |0\rangle \,,
\end{equation}
and the spectrum of gravitational waves ${\cal P}_g(k)$ is then defined by
\begin{equation}
\langle \hat{v}_{{\bf k},\lambda } \hat{v}^*_{{\bf l},\lambda} \rangle
=\frac{m_{{\rm Pl}}^2 a^2}{32\pi}
\frac{2\pi^2}{k^3} {\cal P}_g \delta^{(3)} ({\bf k}-{\bf l}) \,.
\end{equation}
The field equation for $u_k$, derived by varying the action
Eq.~(\ref{rewrite}), is
\begin{equation}
\label{fieldeqn}
\frac{d^2v_k}{d\tau^2} +\left( k^2 -\frac{1}{a}\frac{d^2a}{d\tau^2} \right)
v_k =0 \,,
\end{equation}
and the scale factor term can be written as
\begin{equation}
\frac{1}{a}\frac{d^2a}{d\tau^2} = 2a^2H^2 \left( 1 - \frac{1}{2} \epsilon
\right) \,.
\end{equation}
This puts us in a very similar situation to that for the density
perturbations. The situation is simplified since $a$ appears directly
in the equation of motion rather than $z$, but the strategy is exactly
the same.
For power-law inflation we can again solve exactly by writing
\begin{equation}
\frac{a_{\tau\tau}}{a} =\frac{1}{\tau^2} \left( \mu^2 -\frac{1}{4}
\right) \,,
\end{equation}
where
\begin{equation}
\mu \equiv \frac{3}{2} +\frac{1}{p-1} \,.
\end{equation}
For power-law inflation $\nu$ and $\mu$ coincide, though in general
they do not. The appropriate solution for $v_k$ is given by
Eq.~(\ref{correctform}), as before, after replacing $\nu$ with $\mu$.
It follows, therefore, that
\begin{equation}
\label{tensoramp}
{\cal P}^{1/2}_g (k) =\frac{2}{\sqrt{\pi}} \, 2^{\mu-1/2}
\frac{\Gamma (\mu)}{\Gamma (3/2)} (\mu-1/2 )^{1/2 -\mu}
\left. \frac{H}{m_{{\rm Pl}}} \right|_{k=aH} \,,
\end{equation}
where $P_g$ has been multiplied by a factor of $2$ to account for the
two polarization states. This exact solution was first obtained by
Abbott and Wise (1984a) and we note that for power-law inflation
\begin{equation}
\frac{{\cal P}_g^{1/2}}{{\cal P}_{\cal R}^{1/2}} = \frac{4}{\sqrt{p}} =
4\sqrt{\epsilon} \,.
\end{equation}
The final step is to carry out the expansion in the same way as in the
scalar case to yield the slow-roll expression for the tensor spectrum.
This gives
\begin{equation}
\mu = \frac{3}{2} + \epsilon \,,
\end{equation}
and hence
\begin{equation}
\label{2ndtens}
{\cal P}^{1/2}_g (k) = \left[ 1 - (C+1) \epsilon \right]
\frac{4}{\sqrt{\pi}}
\left. \frac{H}{m_{{\rm Pl}}} \right|_{k=aH} \,.
\end{equation}
\section{~~~Lowest-Order Reconstruction}
\label{first}
\def9.\arabic{equation}{4.\arabic{equation}}
\setcounter{equation}{0}
In the previous section we discussed the derivation of expressions for
the two {\em initial} spectra ${\cal P}^{1/2}_{{\cal{R}}}$ and ${\cal
P}^{1/2}_g$, which were accurate to next-order in the slow-roll
parameters. Before proceeding, let's relate our notation to other
notations that the reader may be familiar with, which concern the
present-day spectra. In order to derive these, one needs the transfer
functions $T(k)$ and $T_g(k)$ for both scalars (Efstathiou, 1990) and
tensors (Turner, White, and Lidsey, 1993) respectively, which describe the
suppression of growth on scale $k$ relative to the infinite wavelength mode.
The transfer functions in general depend on a whole range of cosmological
parameters, as discussed later. The present-day spectrum of density
perturbations, denoted $P(k)$, is given by
\begin{equation}
\frac{k^3}{2\pi^2} P(k) = \left( \frac{k}{aH} \right)^4 \, T^2(k) \,
{\cal P}_{{\cal R}}(k) \,,
\end{equation}
whle the energy density (per octave) in gravitational waves is
\begin{equation}
\Omega_g(k) = \frac{1}{24} \, T_g^2(k) \, {\cal P}_g(k) \,.
\end{equation}
These expressions apply to a critical density universe; for models with a
cosmological constant they require generalization (see Turner and White,
1996). Note though that in the following Sections, we shall always be
working with (rescaled versions of) the initial spectra, and not with the
present-day spectra.
In this Section, we shall concentrate on the lowest-order situation,
where all expressions are truncated at the lowest-order. This is not
equivalent to assuming that $\epsilon$ and $\eta$ are zero, for in
some expressions, such as the spectral indices, the lowest-order terms
contain $\epsilon$ and $\eta$, as we shall see. This approximation
can be regarded as being extremely useful for the present state of
observations. However, optimistically one hopes that future
observations, particularly satellite-based high resolution microwave
background anisotropy observations, will require a higher degree of
accuracy as discussed in Section \ref{second}.
\subsection{The consistency equation and generic predictions of inflation}
In the forthcoming analysis it will prove convenient to work with
rescaled expressions for the spectra
${\cal P}^{1/2}_{{\cal{R}}}$ and ${\cal P}^{1/2}_g$ which
we will use throughout the rest of the paper.
To lowest-order we obtain
\begin{eqnarray}
\label{firstscalar}
A_S(k) & \equiv & 2 {\cal P}^{1/2}_{{\cal{R}}}/5 =
\left. \frac{4}{5} \frac{H^2}{m_{{\rm Pl}}^2|H'|}
\right|_{k=aH} \,, \\
\label{firsttensor}
A_T(k) & \equiv & {\cal
P}^{1/2}_g/10 = \left. \frac{2}{5\sqrt{\pi}} \frac{H}{m_{{\rm Pl}}}
\right|_{k=aH} \,.
\end{eqnarray}
The specific choice of normalizations is arbitrary\footnote{We remark
that these expressions have different prefactors to those contained in
our original papers, CKLL1 and CKLL2; while one normalization is as
valid as any other, the normalizations chosen in those papers were
atypical of the literature. Those used here conform more readily with
the conventions employed in the existing literature and in particular
with the Stewart and Lyth (1993) calculation. In fact, the numerical
difference is only 0.3\%. The ratio of the tensor and scalar
amplitudes is unaffected by this change.}. The above choice ensures
that $A_S$ coincides precisely with the quantity $\delta_H$ as defined
by Liddle and Lyth (1993a, Eq.~(3.6)). This parameter may be viewed as
the density contrast at Hubble-radius-crossing. The normalization for
the tensor spectrum is then chosen so that to lowest-order $\epsilon =
A_T^2/A_S^2$.
During inflation the scalar field slowly rolls down its
self-interaction potential. This causes the Hubble parameter to vary
as a function of cosmic time and therefore with respect to the scale
at Hubble-radius crossing. The expressions for the perturbations
therefore acquire a dependence on scale and it is conventional to
express this variation in terms of {\em spectral indices}. In general,
these indices are themselves functions of scale and there appear to be
two ways in which they may be defined. In the first case, one may
simply write the power spectra as
\begin{equation}
A_S^2(k) = A_S^2(k_0) \left(\frac{k}{k_0}\right)^{\tilde{n}(k)-1} \quad ;
\quad A_T^2(k) =A_T^2(k_0) \left(\frac{k}{k_0}\right)^{\tilde{n}_T(k)} \,.
\end{equation}
Although these definitions are completely general, they do require a
specific choice of $k_0$ to be made. This feature implies that the
definitions are non-local, a considerable drawback. A more suitable
alternative is to define the spectral indices differentially via
\begin{eqnarray}
n(k)-1 & \equiv & \frac{d \ln A_S^2}{d \ln k} \,,\\
n_T(k) & \equiv & \frac{d \ln A_T^2}{d \ln k} \,.
\end{eqnarray}
We shall adopt this second choice in this work. The two definitions
coincide for power-law spectra, where the indices are constant. In
general, however, they are inequivalent.
At the level of approximation we are considering in this Section, the
spectral indices may be expressed directly in terms of the slow-roll
parameters $\epsilon$ and $\eta$. One calculates the first derivatives
of the amplitudes from Eqs.~(\ref{firstscalar}) and
(\ref{firsttensor}) with respect to $\phi$ and converts to derivatives
with respect to wavenumber with the help of Eq.~(\ref{kphi}). It is
straightforward to show that
\begin{eqnarray}
\label{scalind1}
n(k)-1 & = & 2 \eta - 4 \epsilon \,, \\
\label{tensind1}
n_T(k) & = & - 2 \epsilon \,.
\end{eqnarray}
The conventional statement attached to these expressions is that
inflation predicts spectra which to the presently desired accuracy can
be approximated as power-laws; that is, that the slow-roll parameters
can be treated as constants. While this statement is formally correct,
it requires some discussion. In particular it is important to realize
that the power-law approximation has no direct connection to the
slow-roll approximation, but rather is a statement that the relevant
observations cover only a limited range of scale and do so with
limited accuracy. As far as the derivations of the spectra are
concerned, the approximation is that for each scale the parameters can
be treated as constant while that scale crosses outside the Hubble
radius. However, in this `adiabatic' approximation, there is no need
for those constant values to remain the same from scale to scale.
Thus, the expressions for the spectra can be applied across the
complete range of scales. Although they are an approximation at each
scale, the approximation does not deteriorate when one attempts to
study a wider range of scales. The feature that dictates whether the
spectra can be treated as power-laws is that the range of scales over
which observations can be made is quite small, in terms of the range
of $\phi$ values, and taking additional derivatives of the spectra
introduces into the lowest-order result an extra power in the
slow-roll parameters. For example, although differentiating
Eq.~(\ref{scalind1}) gives the correct lowest-order expression for
$dn/d \ln k$, this will be of order $\epsilon^2$ and hence a small
effect over the short range of scales large-scale structure samples.
Were large-scale structure able to sample, for example, scales
encompassing twenty orders of magnitude rather than four, the
approximation by power-law would be liable to break down for typical
inflation models. With high accuracy observations, the power-law
approximation represented by these lowest-order expressions may prove
inadequate even over the short range of accessible scales.
We emphasize that the spectral indices do not have to satisfy the
exact power-law result $n-1 = n_T$ at this level of approximation.
Each spectrum is uniquely specified by its amplitude and spectral
index. The overall amplitude is a free parameter determined by the
normalization of the expansion rate $H$ during inflation (or
equivalently the scalar field potential $V$). On the other hand, the
relative amplitude of the two spectra is given by
\begin{equation}
\label{rat1}
\frac{A_T^2}{A_S^2} = \epsilon \,.
\end{equation}
Thus, there exists a simple relationship between the relative
amplitude and the tensor spectral index:
\begin{equation}
\label{firstconsistency}
n_T = - 2 \frac{A_T^2}{A_S^2} \,.
\end{equation}
This is the lowest-order consistency equation and represents an
extremely distinctive signature of inflationary models. It is
difficult to conceive of such a relation occurring via any other
mechanism for the generation of the spectra.
Since it is possible for the spectra to have different indices, the
assumption that their ratio is fixed can be true only for a limited
range of scales, but the correction enters at a higher order in the
slow-roll parameters.
This expression is often written in a slightly different form in order
to bring the right hand side closer to observations. Since the spectra
can be defined with arbitrary prefactors, they themselves have no
definite significance. The environment in which each spectrum may have
an effect that allows direct comparison is in large angle microwave
background anisotropies. In this case the scalar and tensor
fluctuations each contribute independently to the expected value of
the microwave multipoles, $C_l$ (defined and discussed in more detail
in Section \ref{obs}), and in the approximation where only the
Sachs-Wolfe term is included and perfect matter--domination at last
scattering assumed, this enables one to write the
lowest-order consistency equation as (Liddle and Lyth, 1992, 1993a)
\begin{equation}
\label{firstCl}
\frac{C_l^T}{C_l^S} = -6.2 n_T \,.
\end{equation}
This equation applies for moderate values of $l$ corresponding to
scales that are sufficiently small for the curvature of the last
scattering surface to be negligible and yet are large enough to be
well above the Hubble radius at decoupling\footnote{The exact number
in this relation is sometimes written in different ways. It was first
evaluated exactly as $25(1+48\pi^2/385)/9$ in the scale-invariant
limit by Starobinsky (1985). This is numerically equal to 6.2. There
is no regime where this strictly holds, as corrections from the `Doppler'
peak and from the Universe being not perfectly matter dominated at last
scattering intervene before the asymptote is reached. Other authors evaluate
only part of the expression to approximate it as $2\pi$, or even $6$.
Finally, many authors consider the ratio of contributions to the
quadrupole $l=2$. In this case there is a geometrical correction from
the curvature of the last scattering surface which make the factor
close to $7$.}.
Eqs.~(\ref{scalind1}), (\ref{tensind1}) and (\ref{rat1}) contain all
the information one requires to determine the generic behavior of
inflationary models at this order. Moreover, the current status of
observational data is such that they are sufficient to allow a
reasonable degree of precision to be attained in the study of
large-scale structure and microwave background anisotropies. In the
forthcoming years, however, data quality will inevitably improve and a
higher degree of accuracy in the theoretical calculations will
therefore be required. Indeed, high precision microwave anisotropy
experiments are likely to be the first type of observation demanding
just such an improvement in accuracy.
In the next Section we shall show how these improvements may be
implemented. For the purposes of our present discussion, however,
there are only two input parameters that need to be determined before
one can proceed to investigate inflation-inspired models of structure
formation (Liddle and Lyth, 1993b). The key points are that (a) the
density perturbation spectrum has a power-law form and that (b) some
fraction of the large angle microwave anisotropies might be due to
gravitational waves. These conditions represent two completely
independent parameters, but fortunately, they are the only two new
parameters one requires in the lowest-order approximation. This is
true even though one has complete freedom in choosing the functional
form of the underlying inflationary potential. A large number of
papers have now investigated the implications of these inflationary
parameters for structure formation models such as Cold Dark Matter and
Mixed Dark Matter models. Some only consider the possibility of tilt
(Bond, 1992; Liddle, Lyth, and Sutherland, 1992; Cen et al., 1992, Pogosyan
and Starobinsky, 1995) and some also allow for gravitational waves
(Liddle and Lyth, 1993b; Schaefer and Shafi, 1994; Liddle et al., 1996).
One can classify the generic behavior of all inflationary models
consistent with the lowest-order approximation into six separate
categories, as summarized in Table 1. Each sector is characterized by
the direction of the tilt away from scale invariant density
perturbations and by the relative amplitude of the gravitational
waves. In general, spectra with $n>1$ increase the short-scale power
of the density perturbation spectrum. Such spectra were named {\em
blue} spectra by Mollerach et al. (1994). Conversely, those spectra
with $n<1$ subtract short scale power\footnote{We resist calling them
{\em red} since the usual definition of red spectra is $n < 0$, not $n
< 1$.}. It is a general feature of inflation that $n < 1$ is easier to
produce than $n>1$. The reason for this follows from the definition
Eq.~(\ref{scalind1}) for the scalar spectral index. To lowest-order, a
necessary and sufficient condition for the spectrum to be blue is
simply that $\eta > 2\epsilon$. Since $\epsilon$ is positive by
definition, this condition is not easy to satisfy and this is
particularly so during the final stages of inflation where $\epsilon$
must necessarily begin to approach unity. However, specific inflation
models have been constructed for each possibility, with the exception
of a blue spectrum accompanied by a large gravitational wave
amplitude. This last possibility, while still technically possible, is
particularly hard to realize because it requires a large $\epsilon$
overpowered by a yet larger $\eta$.
\subsection{Reconstructing the potential}
In CKLL1 we developed a framework initiated by Hodges and Blumenthal
(1990) that one might call {\em functional reconstruction}. In this
approach one views the observations as determining the spectra
explicitly as functions of scale. Hodges and Blumenthal (1990)
considered only scalar perturbations, and then Grishchuk and Solokhin
(1991) made an investigation, considering only the tensors, with the
aim of determining the time evolution of the Hubble parameter. In
CKLL1, we provided a unified treatment of both scalars and tensors.
The ultimate aim of such a procedure is to then process the functions
through the differential equations describing the evolution of the
universe during inflation. One thereby determines the potential
driving inflation as a function of the scalar field. If such a
procedure could be carried out exactly, the quantities in the
consistency equation would also be functions of scale.
An important point worth emphasising here is that only by including
the tensors can a full reconstruction be achieved. The scalar
perturbations only determine the potential up to an unknown constant.
As the underlying equations are non-linear, different choices of the
constant lead not just to a rescaling of the potential but to an
entirely new functional form. Thus, there are many potentials which
lead to the same scalar spectrum, and hence no unique reconstruction
of the potential from the scalar spectrum. Any piece of knowledge
concerning the tensors is enough to break this degeneracy.
{}From a practical point of view, one finds that the functional
reconstruction procedure is not very useful, although it does allow
some theoretical insight to be gained. The reason is that {\em exact}
formulae for the amplitudes of the spectra do not exist for an {\em
arbitrary} inflaton potential. Consequently, even though the classical
dynamics of the scalar field can be accounted for exactly, one must
input the information on the spectra using results that depend
directly on the slow-roll expansion. At some level, it is inconsistent
to treat the dynamics exactly and the perturbations approximately, so
formally one should truncate both at the same order of approximation.
Indeed, the next-order calculations we provide in the following
Section show that this joint truncation is indeed preferable. In
general, the next-order correction to the magnitude of the potential
arising from the spectra has an opposite sign and is slightly larger
than the correction to the dynamics. In effect, therefore, an exact
treatment of the dynamics actually leads to a less accurate answer
than that obtained by treating the entire problem to lowest-order in
slow-roll!
We therefore advocate an alternative approach that may be referred to
as {\em perturbative reconstruction}. The fundamental idea behind
perturbative reconstruction follows directly from the fact that the
scalar field must roll sufficiently slowly down its potential if
inflation is to proceed at all. This is important for the following
reason. Typically, the modes that ultimately lead to observational
effects within our universe first crossed the Hubble radius somewhere
between 50 and 60 $e$-foldings before inflation came to an end. (The
precise number of $e$-foldings depends on the final reheating
temperature, but this does not affect the general features of the
argument). During these 10 $e$-foldings of inflationary expansion, the
change in the value of the inflaton field is typically small. In
effect, therefore, the position of the field in the potential would
have remained essentially fixed at some specific value $\phi_0$. It
follows, that cosmological and astrophysical observations can only
yield information regarding this small segment of the potential.
Hence, it is consistent to expand the underlying inflationary
potential as a Taylor series about the point $\phi_0$. The use of such
a procedure to lowest-order was suggested by Turner (1993a), Copeland
et al. (1993a) and CKLL1. Turner (1993b) then included a next-order
term in the potential. The formalism was then developed fully to
next-order in CKLL2, including a next-order term in the derivatives as
well as the potential and outlining the framework for the general
expansion. This framework was recast into a more observationally-based
language by Liddle and Turner (1994) who further discussed the meaning
of the order-by-order expansion.
Perturbative reconstruction can be performed in a controlled way using
the slow-roll expansion order-by-order. The dynamics can be treated to
arbitrary order in this expansion by employing the formalism developed
by Liddle et al. (1994). In contrast, however, the treatment of
perturbations is presently available only to next-order. In this case
there seems no obvious framework by which one can establish an
order-by-order expansion, and even just obtaining terms to one higher
order is a very difficult task.
Modulo questions of convergence, the perturbative reconstruction
procedure successfully encodes functional reconstruction in the sense
that perturbative reconstruction performed to infinite order is
formally equivalent to functional reconstruction. Perturbative
reconstruction can also be rewritten as an expansion in the observed
spectra. The advantage of considering an expansion of this type is
that it indicates exactly how the features in the observed spectra
yield information on the inflationary potential. Such an explicit
account of the observational expansion has not been given before.
Before launching into specific calculation, however, it will be
helpful to identify each observable quantity with some order in the
slow-roll expansion. This may be achieved by considering which
slow-roll parameters occur in the lowest-order term. Thus, one may
employ the lowest-order expressions for the spectra. One sees by
direct differentiation that the information associated with the
accumulation of observables is as follows: $H$ gives $A_T^2$,
$\epsilon$ gives $A_S^2$ and $n_T$, $\eta$ gives $n$ and $dn_T/d \ln
k$, $\xi$ gives $dn/d \ln k$ and $d^2 n_T/d \ln k^2$, and so on. The
key feature is that the tensor spectrum always remains one step above
the scalar one. Furthermore, we shall see that an additional
derivative of the Hubble parameter for each order is required to
obtain a higher order expression for each observable.
We shall now proceed to derive expressions for the potential and its
first two derivatives correct to lowest-order in the slow-roll
expansion. We consider the Taylor series
\begin{equation}
\label{taylorexpansion}
V(\phi) = V(\phi_0) + V'(\phi_0) \Delta \phi + \frac{1}{2} V''(\phi_0)
\Delta\phi^2 + \cdots \,,
\end{equation}
about the point $\phi_0$. At this order, the Hamilton-Jacobi equation
(\ref{H}) reduces to $V(\phi) = 3m_{{\rm Pl}}^2 H^2(\phi)/8\pi$, so
the derivatives in this expansion may be expressed directly in terms
of the slow-roll parameters from Eqs.~(\ref{epsilon}) and (\ref{eta}).
It is only consistent to expand the potential to quadratic order,
because the third derivative will contain terms that are of the same
order as terms that were neglected in the original expressions for the
amplitudes. In other words, the lowest-order expressions do not permit
any higher derivatives to be obtained.
It follows by direct substitution, therefore, that
Eq.~(\ref{taylorexpansion}) may be written as
\begin{equation}
\label{expand}
V(\phi) = \frac{3m_{{\rm Pl}}^2 H_0^2}{8\pi} \left[ 1- (16\pi\epsilon_0
)^{1/2} \frac{\Delta\phi}{m_{{\rm Pl}}} + 4\pi (\epsilon_0 + \eta_0 )
\frac{\left( \Delta\phi \right)^2 }{m_{{\rm Pl}}^2} +{\cal{O}}
\left( \frac{(\Delta\phi)^3}{m_{{\rm Pl}}^3} \right) \right] \,,
\end{equation}
where a subscript $0$ implies that quantities are to be evaluated at
$\phi =\phi_0$. Hence, $H_0$ represents the expansion rate when the
scale corresponding to this value of the scalar field first crossed
the Hubble radius during inflation.
We write the coefficients that arise in this expansion in terms of the
spectra by employing the expressions Eqs.~(\ref{firstscalar}) and
(\ref{firsttensor}) for the amplitudes, the definition
Eq.~(\ref{scalind1}) for the scalar spectral index and the definitions
of the slow-roll parameters. We find that
\begin{eqnarray}
V(\phi_0) & = & \frac{75 m_{{\rm Pl}}^4}{32} A_T^2(k_0) \,, \\
V'(\phi_0) & = & - \frac{75 \sqrt{\pi}}{8} m_{{\rm Pl}}^3
\frac{A_T^3(k_0)}{A_S(k_0)} \,, \\
V''(\phi_0) & = & \frac{25\pi}{4} m_{{\rm Pl}}^2 A_T^2(k_0) \left[ 9
\frac{A_T^2(k_0)}{A_S^2(k_0)} - \frac{3}{2} (1-n_0) \right] \,,
\end{eqnarray}
where $k_0$ is the scale at which the amplitude and spectral indices
are determined and $n_0$ is the scalar spectral index at $k_0$. As
already implied by Eq.~(\ref{scalind1}), if $n$ exceeds one the
potential must be convex ($V''>0$) at the point being probed. However,
$n$ being less than one says nothing definite about convexity or
concavity.
Perturbative reconstruction can be possible even if it ultimately
transpires that the observations necessary to test the consistency
equation non-trivially cannot acquire sufficient accuracy. Similar
work on reconstruction to this level of approximation has been done by
Adams and Freese (1995), Mielke and Schunck (1995) and Mangano, Miele, and
Stornaiolo (1995).
However, it is clear that a determination of the gravitational wave
amplitude on at least one scale is essential for the reconstruction
program to work. Presently, such a quantity has not been directly
determined, but we may nevertheless draw some interesting conclusions
from the above calculation. In particular, there are a number of
limiting cases to Eq.~(\ref{expand}) that are of interest. Firstly,
when $\epsilon = \eta$, Eq.~(\ref{expand}) is the expansion for the
exponential potential $V \propto \exp (-\sqrt{16\pi \epsilon} \,
\phi/m_{{\rm Pl}})$. (Without loss of generality we may perform a
linear translation on the value of the scalar field such that
$\phi_0=0$). Secondly, the potential has the form
\begin{equation}
\label{taylor}
V(\phi) =\Lambda \left[ 1+ 2\pi (n-1) \phi^2/m_{{\rm Pl}}^2 \right] \,,
\end{equation}
in the limiting case where $\epsilon \ll 1$. This class of potentials
produces a negligible amount of gravitational waves, but a tilted
scalar perturbation spectrum. The tilt arises because the curvature of
the potential is significant. The direction of the tilt, as determined
by the sign of $(n-1)$, depends on whether the effective mass of the
inflaton field is real or imaginary.
The dynamics of inflation driven by a potential of the form
Eq.~(\ref{taylor}) for $n>1$ has an interesting property. The kinetic
energy of the inflaton field is determined from $H'(\phi)$ via the
second expression in Eq.~(\ref{field}). As the field rolls down the
potential towards $\phi =0$, $H'$ gradually decreases whilst $H$ tends
towards a positive constant. Hence, the field slows down as it
approaches the minimum, but it loses kinetic energy in such a way that
it can never reach the minimum in a finite time. Hence, the de Sitter
universe is a stable attractor for this model and consequently the
inflationary expansion can never end.
There are two ways of circumventing this difficulty. Firstly, one can
argue that the potential only resembles Eq.~(\ref{taylor}) over the
small region corresponding to cosmological scales. This is rather
unsatisfactory, however, since it requires ad-hoc fine-tuning of the
potential and therefore goes against the overall spirit of inflation.
A much more plausible suggestion is that the first term of
Eq.~(\ref{taylor}) arises because a {\em second} scalar field is being
held captive in a false vacuum state. This is the case, for example,
in Linde's Hybrid Inflation scenario (Linde, 1991, 1994; Copeland et
al., 1994b), and an associated instability can end inflation.
We end this section by quoting formulae appropriate to the situation
where one is given the potential and must calculate the predicted
spectra; in general, one cannot analytically find the $H(\phi)$
corresponding to a given $V(\phi)$. In order to obtain the spectra,
one uses the Friedmann equation Eq.~(\ref{H}) and its derivatives in
combination with the slow-roll approximation. To lowest-order, the
spectral indices were first given by Liddle and Lyth (1992), and are
\begin{eqnarray}
n - 1 & = & - 6 \epsilon_{{\rm V}} + 2 \eta_{{\rm V}} \,, \\
n_T & = & - 2 \epsilon_{{\rm V}} \,,
\end{eqnarray}
where
\begin{equation}
\epsilon_{{\rm V}} = \frac{m_{{\rm Pl}}^2}{16 \pi} \, \left( \frac{V'}{V}
\right)^2 \quad ; \quad \eta_{{\rm V}} = \frac{m_{{\rm Pl}}^2}{8
\pi} \, \frac{V''}{V} \,,
\end{equation}
are slow-roll parameters defined from the potential and differ
slightly from the definitions made in terms of the Hubble parameter
used in the rest of this paper (see Liddle et al. (1994) for more
details). It is also possible to write down next-order expressions
for the spectral indices in terms of the potential (Stewart and Lyth,
1993; Kolb and Vadas, 1994). Expressions such as these
written in terms of the potential only make sense because of the
existence of the inflationary attractor.
\section{~~~Next-Order Reconstruction}
\label{second}
\def9.\arabic{equation}{5.\arabic{equation}}
\setcounter{equation}{0}
The level of accuracy discussed in the previous Section, while
perfectly adequate at present, is unlikely to be sufficient once high
resolution microwave background anisotropy experiments are carried
out. The theoretical benchmark for calculating the radiation power
spectrum from a matter power spectrum has been set at one percent in
order to cope with such observations (Hu et al., 1995). If inflation is
to take advantage of this level of accuracy, it is vital that the
initial power spectrum can be considered to at least a similar level
of accuracy. At the very least, this will require the next-order
expressions for the spectra, which represent the highest level of
accuracy presently achieved.
For many potentials, the next-order corrections may be small, perhaps
smaller than the likely observational errors on the lowest-order
terms. We shall see this in the simulated example later in this paper.
In such a case the next-order calculation is still useful, because it
serves as an estimate of the theoretical error bar on the calculation,
which can be contrasted with the observational error.
We devote this Section to describing the next-order results in detail.
\subsection{The consistency equations}
Let's first consider the next-order version of the lowest-order
consistency equation Eq.~(\ref{firstconsistency}). The best available
calculations of the perturbation spectra are those by Stewart and Lyth
(1993) containing the next-order, which we reviewed extensively in
Section \ref{pert}. To this order, the amplitudes for the scalar and
tensor fluctuations are given by
\begin{eqnarray}
\label{secondscalar}
A_S(k) & = & \frac{4}{5 m_{{\rm Pl}}^2} \left[ 1-(2C+1) \epsilon +
C\eta \right] \left. \frac{H^2}{|H'|} \right|_{k=aH} \,, \\
\label{secondtensor}
A_T(k) & = & \frac{2}{5\sqrt{\pi}} \left[ 1-(C+1 ) \epsilon \right]
\left. \frac{H}{m_{{\rm Pl}}} \right|_{k=aH} \,,
\end{eqnarray}
respectively, where we choose the same normalizations for $A_S$ and
$A_T$ as in Section \ref{first}. We recall that $C \simeq -0.73$ is a
constant. Once again, the right-hand sides of these expressions are to
be evaluated when the scale in question crosses the Hubble radius
during inflation.
Throughout the remainder of this Section we shall be quoting results
that feature a leading term and a correction term, the next-order
term, which is one order higher in the slow-roll parameters. We shall
utilize the symbol ``$\simeq$'' to indicate this level of accuracy.
The correction terms shall be placed in square brackets, so the
lowest-order equations can always be obtained by setting the square
brackets equal to unity, except in Eqs.~(\ref{scalarindex}) and
(\ref{V''result}) where it
needs to be set to zero.
To next-order, the scalar and tensor spectral indices may be expressed
in terms of the first three slow-roll parameters by differentiating
Eqs.~(\ref{secondscalar}) and (\ref{secondtensor}) with respect to
wavenumber $k$ and employing Eq.~(\ref{kphi}). Some straightforward
algebra yields (Stewart and Lyth, 1993)
\begin{eqnarray}
\label{scalarindex}
1-n & \simeq & 4 \epsilon -2 \eta +\left[ 8(C+1) \epsilon^2 -(6+10 C)
\epsilon \eta +2C \xi^2 \right] \,,\\
\label{tensorindex}
n_T & \simeq & -2 \epsilon \left[ 1+(3+2C) \epsilon -2(1+C) \eta \right] \,.
\end{eqnarray}
A very useful relationship may be derived by considering the ratio of
the tensor and scalar amplitudes and replacing the derivative of the
Hubble expansion rate with $\epsilon$. We find that
\begin{equation}
\label{useful}
\epsilon \simeq \frac{A_T^2}{A_S^2} \left[ 1-2C (\epsilon -
\eta ) \right] \,.
\end{equation}
This relationship is the next-order generalization of
Eq.~(\ref{rat1}). It plays a central role in deriving the next-order
expressions for the potential and its first two derivatives in terms
of observables. Moreover, substitution of this expression into
Eq.~(\ref{tensorindex}) implies that
\begin{equation}
n_T \simeq -2 \frac{A^2_T}{A^2_S} \left[ 1+3 \epsilon -2 \eta \right] \,.
\end{equation}
Now, since all the quantities in the square brackets of this
expression are accompanied by a lowest-order prefactor, they may be
converted into observables by applying the lowest-order expressions
Eqs.~(\ref{scalind1}) and (\ref{rat1}). We conclude, therefore, that
\begin{equation}
\label{secondconsistency}
n_T \simeq -2 \frac{A^2_T}{A^2_S} \left[ 1-\frac{A^2_T}{A^2_S} +(1-n)
\right] \,.
\end{equation}
This is the next-order version of the lowest-order consistency
equation $n_T = -2A_T^2/A_S^2$, given first in CKLL2 and translated
into more observational language by Liddle and Turner (1994). It is
interesting to remark that the corrections entering at next-order
depend only on the relative amplitudes of the spectra and on $n$. They
do {\em not} depend on $n_T$ or on any of the derivatives of the
indices, because they can be consistently removed using the
lowest-order version of the same equation. This has an important
consequence that has only been implicit in the literature thus far. We
anticipate that $n$ will be considerably easier to measure than $n_T$.
It is reasonable to suppose, therefore, that if one has enough
observational information to test the lowest-order consistency
equation, one will also have sufficient data to test the next-order
version as well. In other words, the situation where only the
quantities in the lowest-order consistency equation are known is
unlikely to arise. Consequently, one should employ the next-order
consistency equation when testing the inflationary scenario, rather
than the more familiar version given by Eqs.~(\ref{firstconsistency})
or (\ref{firstCl}).
Another new feature of extending the observables to allow
reconstruction at this order is that one has an entirely new
consistency equation, being the lowest-order version of the derivative
of the original consistency equation. One calculates $dn_T/d \ln k$
by differentiating Eq.~(\ref{tensorindex}) with respect to scale $k$
and employing Eqs.~(\ref{epsilon}) and (\ref{kphi}). One finds that
\begin{equation}
\frac{dn_T}{d \ln k} \simeq -4 \epsilon (\epsilon -\eta ) \,.
\end{equation}
Conversion of this expression into observables follows immediately by
substituting in the lowest-order results Eqs.~(\ref{scalind1}) and
(\ref{rat1}), giving
\begin{equation}
\frac{d n_T}{d \ln k} \simeq 2 \frac{A_T^2}{A_S^2} \left(
2 \frac{A^2_T}{A_S^2} +(n-1) \right) \,.
\end{equation}
This equation was derived by Kosowsky and Turner (1995), though they
did not explicitly recognize it as a new consistency equation.
Unfortunately, the observables appearing in the above expression are far
from promising as regards using it.
\subsection{Reconstruction of the potential to next-order}
Now that we have discussed the formalism necessary for calculating the
dynamics and perturbation spectra up to next-order in the slow-roll
expansion, we shall proceed to consider the reconstruction of the
inflationary potential at this improved level of approximation.
We begin by deriving expressions for the potential and its derivatives
directly from the field equation Eq.~(\ref{H}) and the definitions
Eqs.~(\ref{epsilon}) -- (\ref{xi}) for the slow-roll parameters.
Successive differentiation of Eq.~(\ref{H}) with respect to the scalar
field yields the exact relations
\begin{equation}
\label{V}
V= \frac{m_{{\rm Pl}}^2 H^2}{8\pi} (3-\epsilon ) \,,
\end{equation}
\begin{equation}
\label{V'}
V' = -\frac{m_{{\rm Pl}} H^2}{\sqrt{4\pi}} \epsilon^{1/2}
(3 - \eta ) \,,
\end{equation}
\begin{equation}
\label{V''}
V'' = H^2 \left( 3\epsilon +3\eta - ( \eta^2 +\xi^2 ) \right) \,,
\end{equation}
Our immediate aim is to consider these expressions at a single point
$\phi_0$ and rewrite them in terms of observable quantities. The
amplitude of the potential is derived by substituting
Eqs.~(\ref{secondtensor}) and (\ref{useful}) into Eq.~(\ref{V}):
\begin{eqnarray}
\label{V_0}
V(\phi_0) & \simeq & \frac{75 m_{{\rm Pl}}^4}{32} A_T^2(k_0) \left[ 1 +
\left( \frac{5}{3} + 2C \right)
\frac{A_T^2(k_0)}{A_S^2(k_0)} \right] \,, \\
& \simeq & \frac{75 m_{{\rm Pl}}^4}{32} A_T^2(k_0) \left[ 1 +
0.21 \frac{A_T^2(k_0)}{A_S^2(k_0)}\right] \,.
\end{eqnarray}
At this stage, it is interesting to consider how this result would be
altered if one treated the scalar field dynamics in full generality
rather than truncating at next-order. It follows from the general
expression Eq.~(\ref{V}) for the potential that the numerical factor
on the next-order term in the last expression of Eq.~(\ref{V_0}) would
become $-1/3$. What this means is that the next-order correction to
the potential that is due to the spectra dominates the dynamical
corrections. This is true for all inflationary models. Since the sign
of the spectral correction is opposite to that of the dynamical ones,
the overall sign of the correction is reversed.
Since the potential's first derivative contains $\eta$, we need
information regarding the value of the scalar spectral index at $k_0$
if we are to obtain $V'(\phi)$. We replace the $H^2$ term in
Eq.~(\ref{V'}) by substituting the tensor amplitude
Eq.~(\ref{secondtensor}) and collecting together the terms containing
$\{ \epsilon , \eta \}$ to linear order. These may then be written in
terms of the spectra via the lowest-order expressions
Eqs.~(\ref{scalind1}) and (\ref{firstconsistency}). The result is
\begin{eqnarray}
V'(\phi_0) & \simeq & - \frac{75 \sqrt{\pi}}{8} m_{{\rm Pl}}^3 \,
\frac{A_T^3(k_0)}{A_S(k_0)} \left[ 1 + (C+2) \epsilon +
(C-1/3) \eta \right] \,, \nonumber \\
& \simeq & - \frac{75 \sqrt{\pi}}{8} m_{{\rm Pl}}^3 \,
\frac{A_T^3(k_0)}{A_S(k_0)} \left[ 1 + 1.27 \epsilon
-1.06 \eta \right] \,, \nonumber \\
& \simeq & - \frac{75 \sqrt{\pi}}{8} m_{{\rm Pl}}^3 \,
\frac{A_T^3(k_0)}{A_S(k_0)} \left[ 1 -0.85
\frac{A_T^2(k_0)}{A_S^2(k_0)} +0.53 (1-n_0) \right] \,.
\end{eqnarray}
The calculation for $V''(\phi_0)$ is much more involved. A new
observable is needed to determine $\xi$; the easiest example being the
rate of change of the scalar spectral index. This will be
substantially harder to measure, though, and it is fortunate that it
only enters at next-order. (However, it would enter at leading order
in $V'''(\phi_0)$, as mentioned in CKLL2 and derived fully in Liddle
and Turner (1994)). We can obtain the next-order correction to
$V''(\phi_0)$ directly in terms of the slow-roll parameters by
employing Eqs.~(\ref{secondtensor}) and (\ref{V''}). We find that
\begin{equation}
\label{VPPSR}
V''(\phi_0) \simeq \frac{75 \pi}{4} m_{{\rm Pl}}^2 A_T^2(k_0) \left(
\epsilon + \eta \right) \left[ 1 + (2C+2) \epsilon -
\frac{1}{3} \left( \frac{\eta^2 +\xi^2}{\eta +\epsilon}
\right) \right] \,.
\end{equation}
To proceed, we must convert the prefactor $(\epsilon +\eta )$ into
observables, accurate to next-order. To accomplish this we must employ
the next-order result Eq.~(\ref{scalarindex}) for the scalar spectral
index. A straightforward rearrangement of this latter equation yields
\begin{eqnarray}
\epsilon+\eta & \simeq & 3 \epsilon \left[ 1 + \frac{4}{3} (C+1)
\epsilon - \left( \frac{3+5C}{3} \right) \eta + \frac{C}{3}
\frac{\xi^2}{\epsilon} \right] - \frac{1-n_0}{2} \,, \nonumber \\
& \simeq & 3 \frac{A_T^2}{A_S^2} \left[ 1 + \frac{1}{3} (4-2C) \epsilon +
\frac{1}{3} (C-3) \eta + \frac{C}{3} \frac{\xi^2}{\epsilon}
\right] - \frac{1-n_0}{2} \,,
\end{eqnarray}
where the second expression follows after substitution of
Eq.~(\ref{useful}). Substituting this into Eq.~(\ref{VPPSR}) yields
\begin{eqnarray}
\label{***}
V''(\phi_0) & \simeq & \frac{225 \pi}{4} m_{{\rm Pl}}^2
\frac{A_T^4(k_0)}{A_S^2(k_0)} \left[ 1 + \frac{4C+10}{3}
\epsilon + \frac{C-3}{3} \eta + \frac{C}{3}
\frac{\xi^2}{\epsilon} \right] \,, \\
& & - \frac{75 \pi}{8} m_{{\rm Pl}}^2 A_T^2(k_0) \left(1-n_0 \right) \left[
1 + (2C+2) \epsilon \right] - \frac{25\pi}{4} m_{{\rm Pl}}^2 A_T^2
(k_0) (\eta^2+\xi^2) \,, \nonumber
\end{eqnarray}
where the last term is entirely next-order. Note that there are two
lowest-order terms. An interesting case is $\eta= -\epsilon$,
corresponding to $H \propto \phi^{1/2}$, for which the lowest-order
term vanishes identically and the final term of Eq.~(\ref{VPPSR}) is
the only one to contribute. The second derivative of the potential is
the lowest derivative at which it is possible for the expected
lowest-order term to vanish.
The final step is to convert the next-order terms into the
observables. As they are already next-order, one only needs the
lowest-order term in their expansion to complete the conversion. From
the lowest-order expression for the scalar spectral index, one finds
to lowest-order that
\begin{equation}
\label{xin'}
\frac{\xi^2}{\epsilon} \simeq - \frac{1}{2\epsilon} \left.
\frac{dn}{d\ln k} \right|_{k_0} + 5 \eta - 4 \epsilon \,.
\end{equation}
Note that the derivative of the spectral index is of order
$\epsilon^2$. Finally, substitution of Eqs.~(\ref{scalind1}),
(\ref{firstconsistency}) and (\ref{xin'}) into Eq.~(\ref{***}) yields
\begin{eqnarray}
\label{V''result}
V''(\phi_0) & \simeq & \frac{25\pi}{4} m_{{\rm Pl}}^2 A_T^2(k_0) \left\{ 9
\frac{A_T^2(k_0)}{A_S^2(k_0)} - \frac{3}{2} (1-n_0) +
\left[ (36C+2) \frac{A_T^4(k_0)}{A_S^4(k_0)} \right. \right. \nonumber \\
& - & \left. \left. \frac{1}{4} (1-n_0)^2 - (12C-6)
\frac{A_T^2(k_0)}{A_S^2(k_0)} (1-n_0) -\frac{1}{2} (3C-1)
\left. \frac{dn}{d \ln k} \right|_{k_0} \right] \right\} \,,
\end{eqnarray}
where the first two terms in the curly brackets represent the
lowest-order contribution.
Before we conclude this section, it is worth remarking on a point that
has perhaps been implicit in the existing literature but has not been
stated explicitly before. A determination of each successive
derivative of the potential requires an extra piece of observational
information. In particular, for the case of lowest-order perturbative
reconstruction, we conclude that the first term in the Taylor
expansion requires only $A_T$, but the second requires both $A_S$ and
$A_T$. The third term, on the other hand, needs both of these together
with $n_0$. The ability to make the observations therefore dictates
how many derivatives we can determine. On the other hand, a comparison
of the lowest-order and next-order expressions for the derivatives
implies the following: the new piece of information necessary for the
derivation of the lowest-order term in $V'$ is also sufficient to
yield the next-order term in $V$. Likewise, the next observation will
give the lowest-order term in $V''$ and this is enough to give the
next-order term in $V'$. Furthermore, it is also sufficient, {\em in
principle}, to give the third-order term in $V$. We stress {\em in
principle} because the theoretical machinery has not been developed to
allow the calculation of a third-order term in the potential or its
derivatives to be performed. Hence, while observational limitations
constrain how high a derivative we can reach, it may be theoretical
rather than observational limitations which prevent higher accuracy in
the lower derivatives. This will be the case even though the necessary
observational information may become available.
Table 2 lists the inflation parameters required for reconstruction of a
given derivative of the potential. Reconstruction requires the inflation
parameters in terms of observables. Relations between inflationary
parameters and observables are given in Tables 3 and 4. A combination of
information from Table 2 and Table 4 results in Table 5, the observables
needed to reconstruct a given derivative of the potential to a certain
order. Although we know the information required for the next-to-next order
given in Table 5, we don't know the coefficients of the expansion.
\section{~~~The Perturbative Reconstruction Framework}
\label{genfram}
\def9.\arabic{equation}{6.\arabic{equation}}
\setcounter{equation}{0}
Although the next-order results of the previous Section represents the
theoretical state-of-the-art, it is possible to see how the general
pattern goes. We discuss this in this Section and also introduce an
expansion of the observations corresponding to perturbative
reconstruction.
\subsection{A variety of expansions}
During reconstruction, there are three types of expansion being
carried out. There is an expansion in terms of observables, an
expansion in terms of slow-roll parameters and an expansion of the
potential itself.
Since the underlying theme behind the reconstruction program is that
one is driven by observations, let us first consider what information
might be available. The reconstruction program assumes some
measurements of $A_S(k)$ and $A_T(k)$ are available over some range of
scales. In practice, the likely range of observations for the scalars
will probably be no greater than $-5 < \ln(k/k_0) < 5$, with a much
shorter range for the tensors. In accordance with the perturbative
reconstruction strategy, the spectra should be expanded about some
scale $k_0$ which corresponds to the scale at horizon crossing when
$\phi = \phi_0$. The appropriate expansion is in terms of
$\ln(k/k_0)$, and of course it makes best sense to carry out the
expansion about a wavenumber close to the middle of the available
data.
In general, the expansions can be written as
\begin{eqnarray}
\label{obse}
\ln A_S^2(k) & = & \ln A_S^2 (k_0) + (n(k_0)-1) \ln \frac{k}{k_0} +
\frac{1}{2} \left. \frac{dn}{d\ln k} \right|_{k_0} \ln^2
\frac{k}{k_0} + \cdots \,,\\
\ln A_T^2(k) & = & \ln A_T^2 (k_0) + n_T(k_0) \ln \frac{k}{k_0} +
\frac{1}{2}
\left. \frac{dn_T}{d\ln k} \right|_{k_0} \ln^2 \frac{k}{k_0} +
\cdots \,,
\end{eqnarray}
where the coefficients continue as far as the accuracy of observations
permit. There is no obligation for the two series to be the same
length. Indeed, we anticipate that information associated with the
scalars will be considerably easier to obtain in practice.
The range of $\ln k$ over which data are available leads to the range
of $\phi$ over which the reconstruction converges well. Notice that
since we believe $\ln(k/k_0)$ can be somewhat greater than unity,
convergence of this type of series will only occur if the successive
coefficients become smaller. Fortunately, we have already seen in
Section \ref{first} that the lowest-order inflationary predictions
attach an extra slow-roll parameter to each higher derivative of the
spectra taken, so convergence can still occur as long as the slow-roll
parameters are smaller than $1/\max{|\ln(k/k_0)|}$. This forms a good
guide as to how wide a range of scales can be addressed via
perturbative reconstruction. The observation that the spectral index
(at least of the scalars) is not too far from unity suggests that the
slow-roll parameters are small. Hence, the observational expansion
might continue to converge well outside the range of $\ln k$ actually
observed. The equivalent statement regarding the potential would be to
say that if it is reconstructed very smoothly for the range $\Delta
\phi$ corresponding to observations, one should feel fairly confident
in continuing the extrapolation of the potential beyond the region
where direct observations were available (though in a practical sense
this does not correspond to any extra information).
The observational expansion discussed above is closely related to the
slow-roll expansion. In particular, we may consider the expansion of
the spectra at a given $k$ in terms of slow-roll parameters, as
discussed in Section \ref{pert}. A qualitative comparison of the two
expansions then yields a general pattern. Each term from the scalars
allows the determination of one extra slow-roll parameter. With regard
to the tensors, a single piece of information (presumably the
amplitude) is necessary before one can proceed at all, as we have
discussed previously. Beyond that, however, extra terms for the
tensors do not provide new slow-roll parameters. Instead, they lead to
degenerate information and hence consistency relationships. If one
has the first two terms for the tensors and the first scalar term, one
can test the single familiar consistency equation $n_T = - 2
A_T^2/A_S^2$. Further tensor terms result in a whole hierarchy of
consistency equations, as we shall discuss further in the next
Subsection.
By including terms consisting of products of more and more slow-roll
parameters, one builds up a more accurate answer. However, there are
two separate factors that prevent arbitrary accuracy from being
obtained. The first is observational limitations. For a practical
observational data set with error bars, the observational expansion
discussed above can only be carried out to some term, beyond which the
coefficients are determined as being consistent with zero within the
errors. (If the error bars are still small when this happens, it may
still correspond to useful information). This reflects directly on the
number of slow-roll parameters $\epsilon$, $\eta$, $\xi$, etc, that
one can measure. In general, however, there are an infinite number of
slow-roll parameters, and formally they are all of the same order
(meaning that for a `generic' potential, one expects them all to be of
similar size). This appears to be rather problematic, since a finite
number of terms in the observational expansion cannot constrain an
infinite number of slow-roll parameters. Fortunately, however, only a
finite (and usually small) number of such terms ever appear when a
specific expression is considered.
The second restriction is that current technical knowledge concerning
the generation of the spectra, as reviewed in Section \ref{pert}, only
allows the calculation of a lowest-order term plus a correction
involving single slow-roll parameters. In general, one anticipates
further corrections including products of two or more slow-roll
parameters, but that has not been achieved. It follows, therefore,
that the number of derivatives in the potential that may be calculated
is determined by observational restrictions, whilst the accuracy of
each derivative is also constrained by theoretical considerations.
It should be emphasized that once an expression written as an
expansion in slow-roll parameters has been found, it can be
differentiated an arbitrary number of times. It is interesting that
the derivatives are accurate to the same number of orders in the
slow-roll parameters. This follows because differentiation respects
the order-by-order expansion. However, differentiation introduces
higher and higher slow-roll parameters from the infinite hierarchy. An
important point here is that the `lowest-order' can be a product of
any number of slow-roll parameters; the phrase is not synonymous with
setting the slow-roll parameters all to zero.
Having started with the observations, we now come round to the crux of
the reconstruction process: the inflaton potential. In perturbative
reconstruction, one aims to calculate the potential and as many of its
derivatives as possible {\em at a single point} to some level of
accuracy in slow-roll parameters. The ultimate goal is to use this
information to reconstruct some portion of the potential about this
point, by carrying out some expansion of $V(\phi)$ about the point
$\phi_0$. The simplest strategy is to use a Taylor series
\begin{equation}
V(\phi) = V(\phi_0) + V'(\phi_0) \Delta \phi + \frac{1}{2} V''(\phi_0)
\Delta\phi^2 + \cdots \,,
\end{equation}
and we shall only consider that case here. The literature does include
more ambitious strategies such as Pad\'{e} approximants and these may
become useful when specific data are available (Liddle and Turner,
1994). The success of this expansion is governed by how far away from
$\phi_0$ one hopes to go, which ultimately arises from the range of
observations one has available, as well as on how accurately the
individual derivatives are determined.
This expression shows us that perturbative reconstruction of the
potential actually involves {\em two} expansions. We have already seen
that the potential is obtained up to some accuracy in the slow-roll
expansion. However, for reconstruction to be successful, it is also
imperative to consider how accurate the expansion in $\Delta \phi$
might be. Determining the coefficients of only the first one or two
terms may be completely useless if $\Delta \phi$ turns out to be
large.
The key to investigating this is to rewrite $\Delta \phi$ in terms of
$\Delta
\ln k$, the range of scales over which observations can realistically be
expected to cover\footnote{Turner (1993b) and Liddle and Turner (1994)
carried out a similar analysis using $\Delta N$, the number of
$e$-foldings. This is perfectly valid but somewhat harder to
interpret in terms of observable scales since it is only formally
equivalent in a lowest-order approximation. In this work, however, we
desire a simple interpretation of the next-order results.}. Broadly
speaking this corresponds to the interval from $1$ Mpc to about $10^4$
Mpc, so assuming a center point in the middle of this region implies a
range for $\Delta \ln k$ between $\pm 5$. This may be biased through
tensor data only being available on large scales, though it will also
be of considerably lower quality than the scalar data. The
relationship that allows one to achieve the comparison between $\Delta
\phi$ and $\Delta \ln k$ is the exact formula Eq.~(\ref{kphi})
presented earlier
\begin{equation}
\frac{d\phi}{d\ln k} = \frac{m_{{\rm Pl}}^2}{4 \pi} \frac{H'}{H} \,
\frac{1}{\epsilon -1} = \frac{m_{{\rm Pl}}}{\sqrt{4\pi}} \,
\frac{\sqrt{\epsilon}}{\epsilon -1} \,,
\end{equation}
together with its derivatives. One can then expand $\Delta \phi$ in
terms of $\Delta \ln k$, expanding each coefficient up to some order
in the slow-roll expansion. Such an expansion begins
\begin{eqnarray}
\label{phitok}
\Delta \phi & = & - \frac{m_{{\rm Pl}}}{\sqrt{4\pi}} \sqrt{\epsilon}
\left[1+\epsilon + \cdots \right] \Delta \ln k \\ \nonumber
&& + \frac{m_{{\rm Pl}}}{\sqrt{16\pi}} \sqrt{\epsilon}
\left[\epsilon - \eta + \cdots \right] (\Delta \ln k)^2 + \cdots \,,
\end{eqnarray}
where, for illustrative purposes, the first coefficient has been given
to next-order in slow-roll and the second one to lowest-order. The
signs are chosen in accordance with our convention that $V'<0$.
For clarity we shall employ $\beta$ to represent a generic slow-roll
parameter. One can then schematically represent the double expansion
(one in $\Delta \phi$ and one in the slow-roll parameters), as
\begin{eqnarray}
\label{cansee}
\frac{V(\phi)}{A_T^2(k_0)} & \sim & \left[ 1 + \beta + \cdots \right] \,,
\nonumber \\
& + & \beta \Delta \ln k \left[ 1+ \beta + \cdots \right] \left\{ 1 + \beta
+ \beta \Delta \ln k + \cdots \right\} \,, \nonumber \\
& + & \beta^2 (\Delta \ln k)^2 \left[ 1 + \beta + \cdots \right] \left\{
1 + \beta + \beta \Delta \ln k + \cdots \right\} \,,
\end{eqnarray}
where numerical constants have not been displayed. The square brackets
represent the expansion of the potential and its derivatives at
$\phi_0$, while the curly brackets represent the $\Delta \phi$, which
itself is written as an expansion in $\Delta \ln k$ with coefficients
expanded in slow-roll.
For the slow-roll expansion to make sense, we need $\beta \ll 1$. One
can see from the schematic layout of Eq.~(\ref{cansee}) that
convergence of the expansion will fail unless $\beta \Delta \ln k \ll
1$, as successively higher-order terms will otherwise become more and
more important. However, we have agreed that $\Delta \ln k$ itself
need not be small. In regions where it is, it is clear that the best
results are obtained by calculating the low derivatives of the
potential as accurately as possible. In regions where $\Delta \ln k$
is not small, however, it is more fruitful to calculate higher
derivatives.
\subsection{The consistency equation hierarchy}
In the previous Subsection, we stated that there exists an infinite
hierarchy of consistency equations. It is not difficult to see why
such a hierarchy should exist. Even though exact expressions for the
spectra as a function of scale are not presently available, one can
imagine having such expressions, at least in principle. In this case,
one could then write down a consistency equation in the full
functional reconstruction framework that applied over all available
scales. This equation could then be represented in the perturbative
reconstruction framework by performing a Taylor (or similar) expansion
on both sides of it. The perturbative consistency equations could then
be derived by equating the coefficients of the expansions. The key
idea here is that the full functional consistency equation and all its
derivatives must be satisfied at the point about which perturbative
reconstruction is being attempted. The equality of each derivative at
this point, however, represents a separate piece of information.
In Section \ref{first} we presented the consistency equation
Eq.~(\ref{firstconsistency}) for lowest-order perturbative
reconstruction. The connection between the tensor--scalar ratio and
the tensor spectral index was first presented by Liddle and Lyth (1992)
and has been much discussed in the literature. This consistency
equation is simply the (unknown) full functional consistency equation
applied at a single point, and moreover, it is the version of that
equation truncated to lowest-order in slow-roll. Indeed, it does not
require a determination of $n$ and it corresponds to the lowest,
non-trivial truncation of the expansion of the observed spectra.
The next order in slow-roll introduces $n$ and $dn_T/d \ln k$. This
not only supplies enough information to impose a next-order version of
the original consistency equation, but is also enough to impose a
lowest-order version of the {\em derivative} of the consistency
equation. The next-order versions of the original consistency equation
were supplied by CKLL2 and Liddle and Turner (1994) and we discussed
these in Section \ref{second}. We also discussed the lowest-order
version of the derivative of the consistency equation in that Section.
This equation was first given by Kosowsky and Turner (1995).
This pattern continues at all orders in the expansion. One can ask why
this has not been emphasised before. One reason is that until now a
clear understanding has not been established regarding the type of
{\em observational} information that appears at each order in the
expansion. At the same stage that one introduces $n$ in the slow-roll
expansion, one should also introduce the rate of change of the tensor
spectral index. The latter does not provide any new information
regarding the reconstruction, in the same way that $n_T$ did not
provide new information at lowest-order in slow-roll. However, it is
subject to the new consistency equation. Researchers have not paid
attention to the new consistency equation because it requires
$dn_T/d\ln k$ and it seems very unlikely that this could ever be
measured.
This concludes our discussion of the theoretical framework for
perturbative reconstruction. In the following Section, therefore, we
shall discuss whether the observations are likely to reach an adequate
level of sophistication in the foreseeable future and then consider a
worked example that illustrates how the reconstruction programme might
be applied in practice.
\section{~~~Worked Examples of Reconstruction}
\label{obs}
\def9.\arabic{equation}{7.\arabic{equation}}
\setcounter{equation}{0}
\subsection{Prospects for reconstruction}
In this Subsection, we shall consider the long-term prospects for
reconstructing the inflaton potential. It is clear that one must
determine the amplitudes of the {\em primordial} power spectra of
scalar and tensor fluctuations on at least one scale, together with
the slope of the scalar spectrum at that scale. Such information would
provide enough information to reconstruct the potential and its first
two derivatives to lowest--order. However, a measurement of $n_T$ is
also required if one is to test the inflationary hypothesis via the
consistency equation. If such information becomes available at all, it
will probably be {\em after} $A_S$, $A_T$ and $n$ have themselves been
determined, so reconstructing to lowest-order should prove easier to
accomplish than testing the scenario via the consistency equation.
It is convenient to separate the full cosmological parameter space
into two sectors. The first contains the inflationary parameters
essential for reconstructing the potential and testing the consistency
equations. They are
\begin{equation}
\label{inflationparameter}
(A_S, r, n, n_T, \cdots) \,,
\end{equation}
where all are evaluated at $k_0$, and the list extends to as many
derivatives of the spectra as one wishes to consider. The
tensor-scalar ratio $r \equiv 12.4A_T^2 /A_S^2$ is defined so that
$r=1$ corresponds to an equal contribution to large angle microwave
anisotropies from the scalar and tensor fluctuations, as follows from
Eqs.~(\ref{firstconsistency}) and (\ref{firstCl})).
The second set consists of the other cosmological parameters:
\begin{equation}
\label{cosmologicalparameter}
(\Omega_0, \Omega_{\Lambda}, \Omega_{{\rm CDM}}, \Omega_{{\rm HDM}} ,
\Omega_{\rm B} h^2, h , z_R , \ldots ) \,,
\end{equation}
where the $\Omega$ represent the densities in matter of various sorts,
respectively the total matter density, cosmological constant, cold
dark matter, hot dark matter and baryonic matter. Here $z_R$
represents the redshift of recombination; it may be that this single
parameter is adequate or the full ionization history may have to be
taken into account. In the standard cold dark matter (CDM) model these
parameters take the values $(A_S(k_0),0,1,0)$ and $(1, 0, 0.95, 0,
0.0125, 0.5)$ respectively (further parameters concerning derivatives
of the spectral indices in the first set being zero); that is, the
scalar amplitude is the only free parameter available to fit to
observations. The standard ionization history of the universe is also
assumed.
Experiments measuring microwave background anisotropies offer the most
promising route towards acquiring such information to within the
desired level of accuracy. Although redshift surveys provide valuable
insight into the nature of the scalar spectrum at the present epoch,
uncertainties in the mass--to--light ratio of galaxy distributions
imply that it is very difficult to determine the primordial spectrum
from these observations alone. There are further complications
associated with uncertainties in the type of non-baryonic dark matter
in the universe. These can lead to significant modifications in the
form of the transfer function. One crucial advantage that microwave
background experiments have, however, is that the level of anisotropy
above 10 arcmin is almost independent of whether the dark matter is
hot or cold (Seljak and Bertschinger, 1994; Stompor, 1994; Ma and
Bertschinger, 1995; Dodelson, Gates, and Stebbins, 1996). Moreover, as we
shall see in Section \ref{otherways}, a direct detection of the
stochastic background of gravitational waves by laser interferometers
seems highly improbable. Thus, the microwave background anisotropies
appear to be the only practical route at present towards determining
the gravitational wave amplitude.
It is conventional to expand the temperature distribution on the sky
in terms of spherical harmonics
\begin{equation}
\frac{\Delta T}{T_0} =\sum_{l=2}^{\infty}
\sum_{m=-l}^l a_{lm}(r)Y_{lm} ({\bf x}) \,,
\end{equation}
where the monopole and dipole terms have been subtracted out and $T_0
= 2.726$K is the present mean background temperature. The $l$-th
multipole corresponds loosely to an angular scale of $\pi/l$, and a
comoving length scale of $100 h^{-1}$ Mpc at the last scattering
surface subtends an angle of about one degree (for $\Omega_0 = 1$).
Inflation predicts that the $a_{lm}$ are gaussian random variables,
with a rotational invariant expectation value for their variance $C_l
\equiv \langle |a_{lm}|^2 \rangle$. The radiation power spectrum is
defined to be $l(l+1) C_l$; this is exactly constant in the case of a
scale--invariant density perturbation spectrum $(n=1, r=0)$ when the
Sachs--Wolfe effect is the sole source of anisotropy (Sachs and Wolfe,
1967; Bond and Efstathiou, 1987). In general, both tensor and scalar
perturbations contribute to the observed radiation power spectrum, and
for inflation these contributions are independent, so
$C_l=C^S_l+C^T_l$.
Accurate calculations of the $C_l$ from both scalar and tensor modes
require numerical solutions using a Boltzmann code (Bond and Efstathiou,
1987), and this can now be done to an extremely high accuracy, of around one
percent or so (Hu et al., 1995). A recent innovation is a new algorithm
based on an integral solution of the Boltzmann equation (Seljak and
Zaldarriaga, 1996a), which obtains this level of accuracy at much less
computational expense. In principle high quality observations can approach
this accuracy though the question of foreground remains a delicate one (Hu
et al., 1995; Tegmark and Efstathiou, 1996) and so the true observational
accuracy will be less. These types of numerical study seem essential
for high accuracy work, although they are complemented by analytical
approaches, which can be made both for scalars (Hu and Sugiyama, 1995)
and for tensors. The latter case is the easier for two reasons;
firstly, only gravitational effects need to be considered and
secondly, gravitational waves redshift away once they are inside the
Hubble radius, so their main influence is only on the lower
multipoles, up to $l \simeq 100$. Analytic studies, of increasing
sophistication, have been made by Abbott and Wise (1984a, 1984b),
Starobinsky (1985), Turner, White, and Lidsey (1993), Atrio-Barandela and
Silk (1994), Allen and Koranda (1994), Koranda and Allen (1994) and Wang
(1996). These results show good agreement with the numerical
calculations of Crittenden et al. (1993a) and Dodelson, Knox,
and Kolb (1994), who evolve the photon distribution function by applying
first-order perturbation theory to the general relativistic Boltzmann
equation for radiative transfer.
With this calculational power in place, there are two main obstacles
to determining the primordial spectra. These are known as `cosmic
variance' and `cosmic confusion', respectively.
\vspace{.1in}
{\em Cosmic Variance}: A given inflationary model predicts the
quantities $C_l=\langle |a_{lm}|^2 \rangle$, but the observed
multipoles measured from a single point in space are $a_l^2=
\sum_{m=-l}^{+l} |a_{lm}|^2/4\pi$. These only represent a single
realization of the $C_l$. It is well known that a finite sampling of
events generated from a random process leads to an intrinsic
uncertainty in the variance even if the experiment is perfectly
accurate; this is sometimes called sample variance. In the limit of
full sky coverage this uncertainty is known as cosmic variance.
More precisely, the $a_l^2$ are a sum of $2l+1$ Gaussian random
variables and therefore have a probability distribution that is a
$\chi^2$ distribution with $2l+1$ degrees of freedom. Thus, for each
multipole there are $2l+1$ samples, so the uncertainty in the $C_l$ is
given by
\begin{equation}
\label{cosvar}
\frac{\Delta C_l}{C_l} = \sqrt{\frac{2}{2l+1}} \,.
\end{equation}
This implies that cosmic variance is proportional to $l^{-1/2}$ and is
therefore less significant on smaller angular scales. However, for any
given experiment, the beam width limits how high an $l$ can be
obtained before experimental noise intervenes, and anyway in standard
cosmological models the predicted signal cuts off rapidly beyond $l
\sim 1000$ due to the finite thickness of the last scattering surface.
Thus, the information on the tensor components is limited because there is
very little signal in near--scale invariant models for $l \ge 200$ where the
effects of cosmic variance are less significant.
\vspace{.1in}
{\em Cosmic Confusion}: The anisotropy below $l \le 60$ is essentially
determined by the inflationary parameters in
Eq.~(\ref{inflationparameter}), and by $\Omega_0$ and
$\Omega_{\Lambda}$, since it is dominated by the purely gravitational
terms rather than the details of the matter content of the universe.
On the other hand, the anisotropies are highly model dependent for
$l>60$ due to the complexity of the operating physical processes. In
particular, the precise level of anisotropy in this range depends
sensitively on the values of the cosmological parameters listed in
Eq.~(\ref{cosmologicalparameter}). Bond et al. (1994) have suggested
that different sets of values for these parameters sometimes lead to
power spectra which are extremely similar (for a review see Steinhardt,
1994). This leads to degeneracies in determined parameters, which Bond
et al. refer to as `cosmic confusion'. Cosmic confusion is problematic
for the reconstruction program and the degeneracy must be lifted
before it can proceed. Fortunately, things have moved on since the
Bond et al. discussion, and it is now acknowledged that observations
can be carried out at such a high accuracy that the degeneracy is
lifted (Hu et al., 1995, Jungman et al., 1996). Tegmark and Efstathiou
(1996) have found that the microwave background anisotropies can be
determined to very high precision even in the presence of multi-component
foreground noise by the COBRAS/SAMBA satellite.
It should also be noted that other methods are available for
determining cosmological parameters. For example, the primordial light
element abundances imply that $0.009 \le \Omega_{\rm B}h^2 \le 0.022$
and these limits may become stronger as observations of deuterium in
quasar absorption lines improve (Olive et al., 1990; Copi, Schramm, and
Turner, 1995). Furthermore, an accurate measurement of $h$, certainly
to within 10\% , seems achievable with the Hubble Space Telescope
(Freedman et al., 1994), whilst polarization of the microwave
background may provide insight into the ionization history of the
universe (Crittenden, Davis, and Steinhardt, 1993b; Frewin, Polnarev, and
Coles, 1994; Crittenden, Coulson, and Turok, 1994; Kosowsky, 1996). There
has also recently been improved understanding of the possibility of using
polarization to probe gravitational waves (Kamionkowski, Kosowsky, and
Stebbins, 1996; Seljak and Zaldarriaga, 1996b; Zaldarriaga and Seljak,
1996). Because gravitational waves typically contribute more (relative to
density perturbations) to the polarization than to the total anisotropy, and
indeed because one can identify a combination of the polarization
parameters which cannot be induced by density perturbations at all, it may
ultimately be possible to use polarization to do better than the
cosmic-variance limited studies of the temperature alone which we discuss
below.
\vspace{.1in}
In view of this, it is important to consider to what degree the next
generation of satellites will be able to determine the inflationary
parameters in Eq.~(\ref{inflationparameter}). Knox and Turner (1994)
have considered what might be deduced from two experiments $A$ and $B$
whose window functions are centered around $l_A \approx 55$ and $l_B
\approx 200$, respectively. Experiment $B$ only measures anisotropy
due to the scalar fluctuations, whereas $A$ will be sensitive to both
scalar and tensor fluctuations. They considered `standard'
cosmological parameters $h=0.5$, $\Omega_{\rm B} \approx 0.05$,
$\Omega_{\Lambda} =0$ and a scale--invariant spectrum. They concluded
that if the tensor-scalar ratio $r \ge 0.14$, one should be able to
rule out $r=0$ with 95\% confidence 95\% of the time. Thus, the
gravitational wave amplitude should be quantitatively measurable for
$r \ge 0.14$. If $n$ is reduced, the limit is improved slightly to $r
\ge 0.1$. Knox and Turner (1994) further conclude that full--sky
measurements on angular scales $0.5^o$ and $3^o$ should acquire the
sensitivity required for making such a detection.
For reconstruction to proceed at lowest--order, however, one also
requires $C_l^S$ for some $l$ and also the spectral index $n$. Knox
(1995) has simulated a set of microwave background experiments within
the context of chaotic inflation driven by a $\phi^4$ potential. This
model predicts $n=0.94$, $n_T=-0.04$ and $r=0.28$. He considers a
third measurement made on a smaller angular scale than those of $A$
and $B$. It is this measurement that determines $C_l^S$ and this may
be combined with the measurement at the intermediate scale $l_B$ to
determine the slope $n$. Finally, $r$ is inferred by identifying the
`excess power' arising in measurement $A$ with the gravitational
waves. He concludes that the quantity $C_2^S \,130^{1-n}$ could be
measured to an accuracy of $\pm 0.3$\% and the error in the slope of
the scalar spectrum could be as small as $\pm 0.02$. If $n \approx 1$,
the error on $r$ is $\pm 0.1$ and improves slightly for smaller $n$. A
full--sky experiment designed with current technology and with a $20'$
beam should be able to achieve such precision.
However, these results are derived on the assumption that the
cosmological parameters have been accurately determined by other
means. Indeed, to achieve the above precision on $r$ and $n$, one
requires the errors in $\Omega_{\rm B}h^2$ to be no more than $10$ \%
and $6 \% $, respectively (Knox, 1995). Furthermore, the Hubble
parameter will have to be determined to within $6\% $ or $14$\%
respectively if $\Omega_{\Lambda} =0.8$ and the uncertainty in
$\Omega_{\Lambda}$ must be below $7\%$.
More recently, Jungman et al. (1996) have carried out an analysis
where all inflationary and cosmological parameters are allowed to
vary. They confirm the expectation that the estimates provided by Knox
(1995) are very optimistic. If all the other cosmological parameters
are left completely free, it is impossible to get any useful
information on the gravitational waves at all --- the required value of $r$
is somewhat larger than mentioned above, and $n_{{\rm T}}$ would have to be
extremely large. However, that represents a somewhat pessimistic assessment,
because certainly many of the cosmological parameters will be constrained by
other types of observations, and more importantly one may also feel content
to live within a subset of cosmological parameter space (for example,
critical density universes with only cold dark matter).
The accuracy to which the above parameters can be observationally
determined will decide whether the information is good enough to push
any of the expressions beyond lowest-order. Another possibility is
that a more sophisticated observable may become available; Kosowsky and
Turner (1995) have considered the possibility that $dn/d \ln k$ might
be observable in the microwave background. For most models this seems
unlikely as the effect will be small, but there do exist inflationary
models leading to an effect that is large enough to be observable.
Whether this parameter generates any degeneracies with other
inflationary or cosmological parameters in the shape of the $C_l$
remains to be addressed.
\subsection{Toy model reconstructions with simulated data}
We devote this subsection to carrying out a worked example of
reconstruction on a faked data set, to indicate the kind of accuracy
that might be possible. We have tried to make the outcome of analyzing
the simulated data at least reasonably indicative of the sort that
high resolution microwave background experiments might achieve, based
on the analysis by Knox (1995) [see also Jungman et al., 1996].
However, our approach is strictly a toy model; it is not intended to
bear any resemblance to what one might actually do with high accuracy
observations. It seems very unlikely that observations such as CMB
anisotropies might be used to directly estimate the $k$-space spectra
(though such an approach is common with galaxy redshift surveys); the
expectation is that if suitable quality data are obtained then the
appropriate procedure will be to push the theory forward from the
spectra rather than try to calculate the primordial spectra directly
from the observations. That is, some analysis such as a likelihood
analysis would be used to find best fitting parameters such as the
amplitude and spectral indices of the scalars and tensors directly.
Knox (1995) has taken some first steps in this direction.
Perturbative reconstruction requires an expansion of the observations
about a single scale, which will end up corresponding to the location
$\phi_0$ on the potential about which it is to be reconstructed. As
discussed earlier, an expansion of the logarithm of the spectra in
terms of the logarithm of the wavenumber is the best way to proceed.
It will always make the most sense to choose the scale $k_0$ about
which the expansion is done to be near the `central' point of the
logarithmic $k$-interval\footnote{The word `central' is in quotes to
indicate that the effective center point of the data may be biased
through tensors only being available on large scales, plus
scale-dependent error bars on both scalars and tensors. The word is
intended to refer to the point best determined by the data assuming
the type of fit attempted.}. Thus we write
\begin{eqnarray}
\ln A_S^2(k) & = & \ln A_S^2 (k_0) + (n(k_0)-1) \ln \frac{k}{k_0} +
\frac{1}{2} \left. \frac{dn}{d\ln k} \right|_{k_0} \ln^2
\frac{k}{k_0} + \cdots \,,\\
\ln A_T^2(k) & = & \ln A_T^2 (k_0) + n_T(k_0) \ln \frac{k}{k_0} +
\frac{1}{2} \left. \frac{dn_T}{d\ln k} \right|_{k_0} \ln^2
\frac{k}{k_0} + \cdots \,,
\end{eqnarray}
where we have written in explicitly the observational quantities to
which the coefficients of the expansion correspond.
A given observational program produces some finite set of data with
error bars, such as a list of galaxy redshifts and sky positions, or a
pixel map of the microwave sky. As we said above, it is unlikely to be
a useful strategy to try and obtain the power spectra from these, and
then use these to reconstruct. Rather, one should push the theory
towards the data by parametrizing the spectra and fitting for those
parameters, as has been done so successfully with COBE. Other
parameters which affect the data interpretation, such as the
cosmological parameters, can be fixed or simultaneously fitted as
required. The general reconstruction framework we have described
indicates an efficient parametrization of the spectra that could be
used.
Despite the above, for our illustrative examples we have chosen
to simulate data for the spectra themselves, as it is the simplest
thing to do. Enough is known (Knox, 1995; Jungman et al., 1996) about
the capabilities of CMB satellites in particular to enable a fairly
realistic example (in terms of the observational uncertainties) to be
constructed. To do anything else would obscure the principal issues. Our aim
therefore is to simulate a set of data, with errors, for the spectra, which
when fitted give similar errors on parameters to those expected had we
carried out the full task of simulating say a microwave sky and fitting
directly for the spectral parameters. It is well outside the scope of this
paper to attempt a realistic simulation of what future data might actually
look like.
As a simple test, we have simulated fake data sets for two different
models, as follows:
\begin{enumerate}
\item A power-law inflation model with power-law index $p = 21$, chosen
to yield $n-1 = n_T = -0.1$. Since power-law inflation can be solved
exactly we know the precise amplitude of the spectra corresponding to
a given normalization of the spectra, Eqs.~(\ref{scalaramp}) and
(\ref{tensoramp}). This particular model has been advocated by White
et al. (1995) as providing a good fit to the current observational
data.
\item An intermediate inflation model (Barrow and Liddle, 1993), which
gives a scale-invariant spectrum of density perturbations but still
possesses significant gravitational waves. We choose a version where
scalars and tensors contribute equally to COBE (to be precise, their
contributions to the tenth multipole are chosen to be the same). In
this case, a precise calculation of the spectra cannot be made, so we
compromise by using the next-order approximation to generate the
spectra from the underlying model.
\end{enumerate}
These models both have quite substantial gravitational waves. They
have been chosen to be compatible with present observational data,
though they can be regarded as rather extreme cases which maximize the
chance of an accurate reconstruction.
The simulated data are constructed by the following procedure.
\begin{itemize}
\item The overall normalization reproduces the COBE result.
\item The scalar error bars are consistent with cosmic variance limited
microwave anisotropy observations up to $l = 200$ (except that for
simplicity we have modeled the errors by a gaussian rather than the
formally correct $\chi^2_{2l+1}$ distribution). Other cosmological
parameters, which affect the microwave anisotropy spectrum, are
assumed fixed. The COBRAS/SAMBA satellite can go to much higher $l$,
but of course the other cosmological parameters will be uncertain
which limits the estimation of the inflationary parameters. By
stopping at $l = 200$, we find that the accuracy we obtain is similar
to that suggested by Jungman et al. (1996) for the full problem, so it
serves as a reasonable compromise.
\item For the tensors, reasonable {\it a priori} estimates for the
error bars are harder to establish. We have assumed data corresponding
to $l$ up to 40, which is where the tensor contribution to $C_l$
begins to cut off, and we have chosen error bars so as to reproduce
the observational uncertainty in the tensor amplitude suggested by
Knox (1995). We then accept whatever uncertainty in the tensor
spectral index this gives us, and it happens to be in reasonable
agreement with that suggested by Knox.
\end{itemize}
The simulated data for Model 1 are shown in Figure 2, along with the
best fit reconstructions. Since scalar data runs from $l = 2$ to
$200$, it covers two orders of magnitude in wavenumber, corresponding
to $\Delta \ln k \simeq 4.6$. The input and output parameters are
shown in Table 4. We performed two fits, the first being a power-law
fit and the second also allowing for a variation in the scalar
spectral index (though in fact the underlying spectrum has none). The
Figures and subsequent discussion use the former.
The results for Models 1 and 2 contain no particular surprises.
Although this is intended only to be indicative and certainly falls way
short of the sophistication that can be brought into play on realistic data,
the error bars are probably fairly reasonable. As expected, the tensor
spectral index is the real stumbling block, but at least with these
models one obtains a strong handle on $A_T^2$, thus allowing a unique
reconstruction. For these reconstructions, we find that the lowest-order
consistency equation Eq.~(\ref{firstconsistency}) is indeed satisfied
\begin{equation}
0.108 \pm 0.013 = 2 \frac{A_T^2}{A_S^2} = - n_T = 0.25 \pm 0.10 \,,
\end{equation}
for Model 1 and
\begin{equation}
0.14 \pm 0.02= 2 \frac{A_T^2}{A_S^2} = - n_T = 0.12 \pm 0.11 \,,
\end{equation}
for Model 2. The same is true for
the next-order version Eq.~(\ref{secondconsistency}). For Model 1 we obtain
\begin{equation}
0.114 \pm 0.014 = 2 \frac{A_T^2}{A_S^2}\left[1-\frac{A_T^2}{A_S^2} + (1-n)
\right] = - n_T = 0.25 \pm 0.10 \,,
\end{equation}
whereas for Model 2 we find
\begin{equation}
0.13 \pm 0.02 = 2 \frac{A_T^2}{A_S^2}\left[1-\frac{A_T^2}{A_S^2} + (1-n)
\right] = - n_T = 0.12 \pm 0.11 \,.
\end{equation}
While encouraging, we see that the test is not particularly strong due
to the poorly determined $n_T$. In models where the tensors are even
weaker than considered here, the task of testing the consistency
equation will be yet harder.
Proceeding on to the reconstruction, Table 5 shows lowest-order and
next-order reconstructions, in comparison to the exact underlying
potential for both Models. The consistency equation has been used to
eliminate $n_T$ as it is the most poorly determined quantity. A next-order
version of $V''(\phi_0)$ cannot be obtained without a value for $dn/d\ln
k|_{k_0}$, though the size of the correction could be bounded from the
error bars on the null result. The reconstructed potentials, both
lowest-order and next-order, for Model 1 are shown in Figure 3 in
comparison to the underlying potential. A Taylor series has been used
to generate them, and the range of $\phi$ shown corresponds to the
range of observational data (a range of two orders of magnitude in
$k$) determined using Eq.~(\ref{phitok}).
We see that in both models the lowest-order reconstruction has been very
successful. The errors are dominated by those in measuring the tensor
amplitude. However, in neither case does the next-order result offer a
significant improvement, given the observational error bars. The main
importance of the next-order result appears therefore to be in bounding the
theoretical error, rather than in providing improved accuracy in the overall
reconstruction.
Figure 3 can be compared to a similar figure in Liddle and Turner
(1994), who investigated reconstruction of a similar exponential
potential. However, they did not include any observational errors,
concentrating instead on the theoretical errors and on the efficacy of
different expansion techniques for the potential. They also assumed
reconstruction over a wider range of scales, and had somewhat poorer
convergence of the reconstructed potential through expanding about one
end of the data (the quadrupole) rather than the center.
\section{~~~Other Ways to Constrain the Potential}
\label{otherways}
\def9.\arabic{equation}{8.\arabic{equation}}
\setcounter{equation}{0}
Up until now we have concentrated, at least implicitly, on
observations connected to large-scale structure in the universe,
including microwave background anisotropies. These certainly provide
the best source of constraints on the inflationary potential, and one
should be very pleased at the prospect of obtaining such constraints.
However, they do cover only a small portion of the full inflationary
potential. There is of course no way of uncovering information about
the potential relevant to larger scales (beyond waiting the relevant
number of Hubble times!), but in principle there are a variety of ways
of constraining the potential appropriate to smaller scales. We shall
discuss such possibilities in this Section. In particular, one may
constrain the potential from the fact that inflation must come to an
end some 50 $e$-foldings after the large-scale structure scales pass
outside the Hubble radius. Further constraints are associated with the
scalar and tensor perturbations on small scales. In principle, laser
interferometers could observe the tensor spectrum as a stochastic
background, though we shall see that this is not promising. The
possible overproduction of primordial black holes (PBHs) immediately
after inflation places upper limits on the amplitude of the last
scalar fluctuation to cross the Hubble radius just before inflation
ends, while distortions to the microwave background spectrum limit
scalar fluctuations on mass scales well below large-scale structure
scales.
\subsection{To the end of inflation and the area law}
In traditional inflation models, inflation can come to an end in one
of two ways. The first is via some drastic event, such as a quantum
tunneling (for example in extended inflation) or a sudden instability
(probably connected to a second field, as in hybrid inflation). If
this happened, probably little information can be drawn from the
behavior approaching the end of inflation. The second way inflation
may come to an end is simply by the potential becoming too
(logarithmically) steep to sustain inflation any longer, as in generic
chaotic inflation models, so that $\epsilon$ reaches unity.
Let us see what one can conclude in the latter case. For definiteness,
let us assume that 50 $e$-foldings are supposed to occur after the
scale $k_0$, about which reconstruction is attempted, leaves the
horizon. The modest dependence of this number on the details of
reheating will not be important. By assumption, inflation will end
precisely when $\epsilon = 1$. The number of $e$-foldings which occur
between two scalar field values is given exactly by
\begin{equation}
N = \sqrt{\frac{4\pi}{m^2_{{\rm Pl}}}} \int_{\phi_1}^{\phi_2}
\frac{1}{\sqrt{\epsilon(\phi)}} \, d\phi \,.
\end{equation}
For our purposes, this can be neatly written as an integral constraint
(Liddle, 1994a)
\begin{equation}
\int_{\phi_0}^{\phi_{{\rm end}}} \frac{1}{\sqrt{\epsilon(\phi)}} \,
\frac{d\phi}{m_{{\rm Pl}}} = \frac{50}{\sqrt{4\pi}} \,.
\end{equation}
This can most easily be thought of graphically. We have reconstructed
the value of $\epsilon$ and its derivative at $\phi_0$, and know
$\epsilon(\phi_{{\rm end}}) = 1$. As shown in Figure 4, if we plot the
curve of $1/\sqrt{\epsilon}$ against $\phi/m_{{\rm Pl}}$, it must be
such that it reaches unity just as the area under it reaches
$50/\sqrt{4\pi}$. While there remain many ways in which the curve may
do this, it does exclude some possibilities such as a sudden
flattening of the potential after observable scales leave the
horizon.\footnote{It appears that this can be used to derive an upper
limit, albeit a weak one, on $(\phi_{{\rm end}} - \phi_0)$, from the
knowledge that $\epsilon \leq 1$. In fact this is not the case, since
$H$ starts to exhibit strong variation when $\epsilon$ approaches one.
The number of $e$-foldings should then strictly be characterized by
the increase in $aH$ rather than $a$ alone (see Liddle et al. (1994)
for details). In principle, a yet weaker constraint may be derived by
using energy scale arguments to limit how much $H$ can decrease in the
late stages of inflation, but such a constraint seems too weak to be
worth pursuing.}
\subsection{Local detection of primordial gravitational waves}
A number of authors have examined the possibility that the stochastic
background of primordial gravitational waves produced during inflation
could be detected locally (Allen, 1988; Grishchuk, 1989; Sahni, 1990;
Souradeep and Sahni, 1992; White, 1992; Turner et al., 1993; Liddle, 1994b;
Bar-Kana, 1994). In general, the wavenumber of the gravitational waves
is related to the value of the inflaton field during inflation via the
relation $\ln (k /k_0 ) = 60 - N$, where $N$ is the number of
$e$-foldings before the end of inflation and $k_0 = a_0 H_0 \approx 3
\times 10^{-18} h$ Hz is the wavenumber of the mode that is just
reentering the Hubble radius at the present epoch. Thus, the modes
with wavenumbers associated with the maximum sensitivity of typical
beam-in-space experiments $(\sim 10^{-3} {\rm Hz})$ first crossed the
Hubble radius approximately 25 $e$-foldings before the end of
inflation. A direct detection of such waves would therefore provide
unique insight into a region of the inflationary potential that cannot
be probed by large-scale structure observations. However, we shall see
that this is unlikely to be possible.
There are a number of gravitational wave detectors currently under
construction or proposal (see e.~g.~Thorne, 1987, 1995). The
ground-based Laser Interferometer Gravitational Wave Observatory
(LIGO) should have a peak sensitivity of $\Omega_g \approx 10^{-11}
h^{-2}$ at 10 Hz (Christensen, 1992), where $\Omega_g$ is the energy
density per logarithmic frequency interval. The proposed space-based
interferometers, the Laser Gravitational Wave Observatory in Space
(Faller et al., 1985; Stebbins et al., 1989) and the Laser
Interferometer Space Antenna (Danzmann, 1995) probe lower
frequencies, but with a sensitivity to flat spectrum stochastic
sources which is less than that of LIGO.
After inflation, the evolution of the gravitational wave perturbation
is determined by Eq.~(\ref{fieldeqn}). We have already studied the
effect of modes which have wavelengths greater than the Hubble radius
by the time of last scattering, which contribute to microwave
background anisotropies. However, the scales which can be detected
locally will have re-entered the Hubble radius before the onset of
matter domination. In this regime they behave as radiation, so their
energy density stays fixed during the radiation era but falls during
the matter era. This suppression factor is directly measured by the
radiation density today, $\Omega_{\rm rad} = 4 \times 10^{-5} h^{-2}$.
Thus the predicted amplitude on scales re-entering before
matter-radiation equality is (Allen, 1988; Sahni, 1990; Liddle, 1994b)
\begin{equation}
\Omega_g h^2 = \frac{2}{3\pi} \left( \frac{H}{m_{{\rm Pl}}} \right)^2
\times 4 \times 10^{-5} \,.
\end{equation}
For the inflation models we have been discussing, $H$ always decreases
with time, and hence the primordial amplitude on short scales is
always less than that on large scales\footnote{`Superinflation' models
have been considered within the context of superstring motivated
cosmologies, and it appears that in that case the gravitational wave
amplitude could rise sufficiently on short scales to be detectable
(Brustein et al., 1995). However, no complete model, demonstrating how
superinflation might successfully end, has been constructed thus far
(Brustein and Veneziano, 1994; Levin, 1995a).}. The quadrupole anisotropy
already places an extremely stringent limit on the amplitude of the
spectrum at large scales, and this immediately translates into a
conservative, but robust, constraint across all short scales of
(Liddle, 1994b)
\begin{equation}
\Omega_g h^2 \leq 4 \times 10^{-15} \,.
\end{equation}
This puts the inflationary signal well out of reach of any of the
proposed experiments.
\subsection{Primordial black holes}
It has been conjectured that primordial black holes (PBHs) may form
during the reheating phase immediately after inflation (Khlopov,
Malomed, and Zel'dovich, 1985; Carr and Lidsey, 1993; Carr, Gilbert, and
Lidsey, 1994; Randall, Solja\u{c}i\'{c}, and Guth, 1996; Garc\'{\i}a-Bellido
et
al., 1996). While there are considerable theoretical uncertainties
attached to this possibility, if such formation does occur, it can
constrain the scalar spectrum at very short scales. During inflation
the first scales to leave the Hubble radius are the last to come back
in and this implies that the very last fluctuation to leave will be
the first to return. In some regions of the post-inflationary
universe, the fluctuation will be so large that one expects that the
collapse of a local region into a black hole will become inevitable.
The higher the rms amplitude the larger the fraction of the universe
forming PBHs. The observational consequences of the evaporation of
these black holes then leads to upper limits on the number that may
form and hence on the magnitude of the spectrum on the relevant
scales. Thus, one may constrain the amplitude of the density spectrum
on scales many orders of magnitude smaller than those probed by
large-scale structure observations and microwave background
experiments. These constraints lead to an upper limit on the spectral
index and may therefore provide insight into features of the
inflationary potential towards the end of inflation.
We parametrize the density spectrum in terms of the mass scale $M$
associated with the Hubble radius when a given mode reenters. Hence,
$\delta(M) \propto M^{(1-n)/6}$ defines the scalar spectral index.
PBHs are never produced in sufficient numbers to be interesting if
$n<1$, but they could be if the spectrum is `blue' with $n>1$.
When an overdense region with equation of state $p=\gamma \rho$ stops
expanding, it must have a size greater than $\sqrt{\gamma}$ times the
horizon size in order to collapse against the pressure. The
probability of a region of mass $M$ forming a PBH is (Carr, 1975)
\begin{equation}
\label{beta0}
\beta (M) \approx \delta (M) \exp \left( -\frac{\gamma^2}{2\delta^2 (M)}
\right) \,.
\end{equation}
The constraints on $\beta (M)$ in the range $10^{10}{\rm g} \le M\le
10^{17}{\rm g}$ have been summarized by Carr and Lidsey (1993). In
particular, PBHs with an initial mass $\sim 10^{15}$g will be
evaporating at the present epoch and may therefore contribute
appreciably to the observed gamma-ray and cosmic-ray spectra at 100
MeV (MacGibbon and Carr, 1991). On the other hand, $10^{10} $g PBHs
have a lifetime $\sim$ 1 sec and, if produced in sufficient numbers,
would lead to the photodissociation of deuterium immediately after the
nucleosynthesis era (Lindley, 1980). PBHs of mass slightly below
$10^{10}$g could alter the photon--to--baryon ratio just prior to
nucleosynthesis. An upper limit therefore arises by requiring that
evaporating PBHs do not generate a photon--to--baryon ratio exceeding
the current value $S_0 = 10^9$ (Zel'dovich and Starobinsky, 1976).
Carr et al. (1994) have considered the constraints on $\beta (M)$
below $10^{10}$g. In this region the strongest constraint arises if
evaporating PBHs leave behind stable Planck mass relics (MacGibbon,
1987; Barrow, Copeland, and Liddle, 1992). The observational constraint
from the relics derives from the fact that they cannot have more than
the critical density at the present epoch, $\Omega_{\rm rel} <1$.
The upshot of this analysis is that the spectral index is typically
constrained to be less than about 1.5, depending weakly on assumptions
as to the reheat temperature after inflation and whether one takes
into account the black hole relic constraint. Because the constraint
applies at the end of inflation, on scales greatly separated from the
microwave anisotropies, it is independent of the COBE normalization
and also of the choice of dark matter. However, in this form it
relies on the spectral index being constant right across those scales
(which it would be in the hybrid inflation model (Copeland et al.,
1994b)). For general inflation models it should be reinterpreted as a
specific constraint on the amplitude at the short scales being
sampled.
Finally, a constraint on the amplitude of the spectrum at a scale
corresponding to an horizon mass $\approx 0.1M_{\odot}$ can in principle
be derived from the recent observations of massive compact halo objects
(MACHOs) (Alcock et al., 1993; Aubourg et al., 1993). The estimated mass
range
of these objects suggests that they constitute about 0.1 per cent of the
critical density. Although the favored explanation for these microlensing
events is that they are due to substellar baryonic brown dwarfs, it
is quite possible that MACHOs may be primordial black holes and therefore
non--baryonic in nature (Nasel'skii and Polnarev; Ivanov, Nasel'skii,
and Novikov, 1994; Yokoyama, 1995). Such PBHs could form from vacuum
fluctuations in the manner discussed above if the amplitude of
spectrum is sufficiently high on the appropriate scale. This may be
possible, for example, if the potential has a suitable form (Ivanov et al.,
1994). Alternatively, a spike may be imposed on the underlying spectrum by
the quantum fluctuations of a second scalar field (Yokoyama, 1995; Randall
et al., 1996; Garc\'{\i}a-Bellido et al., 1996). If the amplitude is too
high on this particular scale, however, it would lead to the overproduction
of MACHO-PBHs. Consistency with the observations therefore constrains both
the spectrum and the inflationary potential.
\subsection{Spectral distortions}
A further constraint on $\delta (M)$ over mass scales considerably
smaller than those corresponding to large-scale structure may be
derived by considering departures of the microwave spectrum away from
a pure blackbody. (For detailed reviews see e.~g.~Danese and de Zotti
(1977) and Sunyaev and Zel'dovich (1980)). Above a redshift of $z_y
\approx 2.2\times 10^4 \left( \Omega_{\rm B} h^2 \right)^{-1/2}$,
Compton scattering is able to establish local thermodynamic
equilibrium whenever there is a sudden redistribution or release of
energy into the universe (Burigana, Danese, and de Zotti, 1991). This
produces a Bose--Einstein spectrum $n \propto \left[ \exp (x + \mu )
-1 \right]^{-1}$ that is characterized by a chemical potential $\mu$,
where $x = h \nu /k T$. (A Planck spectrum corresponds to $\mu =0$).
On the other hand, equilibrium cannot be established for redshifts
just below $z_y$. The distribution of energy at this time could
therefore lead to observable spectral distortions $(\mu \ne 0)$ in the
microwave background at the present epoch. The Far Infrared Absolute
Spectrophotometer (FIRAS) aboard COBE has constrained the spectral
distortion to be $|\mu | < 3.3 \times 10^{-4}$ (Mather et al., 1994),
whilst Hu, Scott, and Silk (1994) have strengthened this limit by
considering the COBE measurement of temperature fluctuations on
$10^{\circ}$ (Bennett et al., 1994). They find that $\mu < 5.0
\times 10^5 (\Delta T/T)^2_{10^{\circ}} \approx 6.3 \times 10^{-5}$.
These limits imply that photon diffusion would have been the dominant
mechanism for producing spectral distortions (Daly, 1991). Silk (1967)
first showed that the damping of adiabatic fluctuations can proceed if
their mass scales are below a characteristic mass known as the Silk
mass. At sufficiently early times, the photons and baryons in the
universe are strongly coupled through Thomson scattering and they
therefore behave as a single viscous fluid. When adiabatic
fluctuations reenter the Hubble radius, they set up pressure gradients
and these result in pressure waves that oscillate as sound waves. As
the epoch of recombination approaches, however, the mean--free--path
of the photons increases and the photons are able to diffuse out of
the overdense regions into underdense regions. Thus, the
inhomogeneities in the photon--baryon fluid are damped. The energy
stored in the fluctuations is redistributed by the diffusion of
photons and it is this transfer of energy during the epoch near to
$z_y$ that produces the spectral distortions. The fluctuations that
lead to these potentially observable effects have mass scales in the
range $10^{-3} < M/ {\rm M}_{\odot} < 10^3$ (Sunyaev and Zel'dovich,
1970; Barrow and Coles, 1991).
The observational upper limit on $\mu$ implies an upper limit on the
amplitude of the pressure wave and therefore a limit on $\delta (M)$.
The energy density in a linear sound wave is $\rho u^2$, where $u
\approx c/\sqrt{3}$ is the sound speed. Thus, the dimensionless energy
release caused by the damping is $q \approx \delta^2/3$. It can be
shown that the spectral distortion is given by $\mu \approx 1.4 q$ and
it follows, therefore, that $\delta < 1.46 \sqrt{\mu} \approx 0.01$.
By normalizing the spectrum at COBE scales $(\sim 10^{22}M_{\odot})$,
an upper limit on the spectral index may be derived. Barrow and Coles
(1991) and Daly (1991) assume that the distortion is entirely due to
the largest amplitude wave and deduce a limit of $n< 1.8$ for $M \sim
10^{-3} {\rm M}_{\odot}$. (The limit becomes weaker for larger
scales). Hu et al. (1994) have derived a stronger constraint of
$n<1.5$ by refining these calculations. This is comparable to the PBH
constraints we have just discussed (though somewhat weaker if one
believes the PBH relic constraint). However, it is probably more
reliable because it is based on physics that is relatively well
understood and requires a less severe extrapolation to smaller scales.
\section{~~~Conclusions}
\label{conc}
\def9.\arabic{equation}{9.\arabic{equation}}
\setcounter{equation}{0}
In this paper, we have reviewed the relationship between observations
of microwave aniso\-tropies and of large-scale structure and the
possibility of connecting them to the potential energy of a scalar
field driving inflation. We have argued that, given suitable quality
observations, the inflationary idea can be tested and then features of
the inflationary potential can be directly measured. In many ways this
is remarkable, given that it is impossible, by many orders of
magnitude, for an Earth-based accelerator to pursue this task.
It is predicted that inflation produces both gravitational waves and
density perturbations. Consequently, the employment of observations
may be divided into two main parts. The most challenging is the test
of the inflationary consistency relations; if these prove testable and
are confirmed, it will provide a powerful vindication of the
chaotic inflation paradigm. One could then feel confident in following the
less observationally challenging task of employing observations to
discern information regarding the inflationary potential, in the form
of its value and that of its first few derivatives at a single point.
We have indicated the different approximation schemes that must be
invoked. Of paramount importance is the slow-roll expansion, but this
must also be coupled to an expansion of the observables. In the
simplest instance this latter expansion corresponds to the
approximation of power-law spectra. The lowest levels of approximation
are certainly able to cope with present-day observations of both
microwave anisotropies and large-scale structure. However, in this
work we have been forward looking, since the demands that will be
imposed on theoretical accuracy by future observations, especially
satellite-based microwave background anisotropy measurements, will be
high. Indeed, they could in principle threaten the limits of
present-day theoretical knowledge regarding the calculation of the
spectra.
We must emphasize that our calculations have all been implemented
within the standard paradigm for chaotic inflation. The vast majority
of known viable models can be expressed within this class, either
trivially or by cunning manipulation, but one should bear in mind that
there exist some models of inflation for which this is not the case.
In some examples, such as old versions of the open inflationary
scenario or some multi-field theories, this is because the predictions
turn out to be dependent on initial conditions. Although such a
situation would be unfortunate it is not logically excluded. Other
theories, such as the recently investigated single-bubble open
inflationary models, rely on dynamics that are much more complicated
than that of the standard scenario (Sasaki et al., 1993; Bucher et al.,
1994; Linde, 1995). They therefore lead to a more complicated
relationship between theory and observations. Furthermore, even if the
inflationary hypothesis is indeed correct, it may be the case that the
actual model produces a very low amplitude of gravitational waves (Lyth
1996). This would make them impossible to measure and such a situation would
remove the ability to make a consistency check and thus eliminate most
of the potential for reconstruction.
Finally, there remains every possibility that the entire inflationary idea
is incorrect; if so, one can at least hope that this is manifested in a
failure of the consistency relations. However, it may not prove
possible to test the consistency relations; might one then blunder into
reconstructing a non-existent object? With sufficiently good observations,
such as a CMB satellite will provide, the answer should be no. The $C_l$
spectrum, when it is observed, will contain huge amounts of degenerate
information. If the correct underlying theory is topological defects, (see
for example Vilenkin and Shellard, 1994), the spectral shape should be very
different to any simple inflation model for any values of the cosmological
parameters. One can certainly reconstruct a `potential' which would give the
observed $C_l$, but it would probably be of such a complex form as to have
little particle physics motivation for it, leaving people to search for
other explanations.
In a standard inflation scenario, the $C_l$ give a complete description of
the gaussian perturbations generated. This prediction can also be tested
against the observations; present observations are compatible with
gaussianity though they are not strong enough to give a convincing test. In
the future we can expect such tests to be widely applied. While in principle
it is possible to construct inflation models giving non-gaussian
perturbations, in practice such models are so contrived that again, were
such features detected, one would quickly be looking for a more plausible
theory for the origin of perturbations. It might well also be that the shape
of the power spectrum might be incompatible with the non-gaussian nature,
within the general context of inflation.
The bulk of this review has covered work already discussed in the
literature. We have given an extensive account of the Stewart and Lyth
(1993) calculation of the perturbation spectra, which provides the
accuracy needed to discuss anticipated observations. The
reconstruction framework has then been described to an accuracy which
ought to be sufficient for years to come. However, as well as the
review material, we have brought to light a few new results and
viewpoints and we summarize these here.
\begin{itemize}
\item The consistency equation discussed in the present literature is just
one of an infinite hierarchy of consistency equations, each of which
can be taken (in principle) to arbitrary accuracy in the slow-roll
expansion. Kosowsky and Turner (1995) have written down the form for
the second member and we have reproduced it here. However, it is
probable that only the first consistency equation will ever be tested.
\item We have indicated that since scalar perturbations are much easier to
measure than tensor ones, the appropriate form of the first
consistency equation to consider is not the lowest-order version, but
rather the next-order version. One requires $n_T$ to test the
lowest-order version and it is very unlikely that such observations
would be available without there also being the appropriate ones to
include the next-order version as well. (The only new ingredient in
the next-order version over and above those quantities in the
lowest-order version is $n$).
\item We have been more explicit than previous work as to how observations
of the primordial spectra should be handled in terms of an expansion in
$\ln k$. We discussed how this expansion relates to the slow-roll
expansion. A worked example on simulated data has illustrated these
ideas in action.
\end{itemize}
In conclusion, therefore, the relationship between inflationary
cosmology and large-scale structure observations is well understood
and the theoretical machinery necessary for taking advantage of high
accuracy observations is now in place. These promise the possibility
of constraining physics at energies inaccessible to any other form of
experiment. Such observations are eagerly awaited.
\section*{Acknowledgments}
JEL is supported by the Particle Physics and Astronomy Research
Council (UK). ARL is supported by the Royal Society. EWK and MA are
supported at Fermilab by the DOE and NASA under Grant NAG 5--2788. TB
is supported by JNICT (Portugal). We are grateful for many helpful
discussions with John Barrow, Robert Caldwell, Bernard Carr, Scott
Dodelson, John Gilbert, Martin Hendry, Lloyd Knox, David Lyth, Douglas
Scott, Paul Steinhardt, Reza Tavakol, Michael Turner and Martin White.
\frenchspacing
\section*{References}
\begin{description}
\item Abbott, L. F. and M. B. Wise, 1984a, Nucl. Phys. B {\bf 244}, 541.
\item Abbott, L. F. and M. B. Wise, 1984b, Phys. Lett. B {\bf 135}, 279.
\item Adams, F. C., J. R. Bond, K. Freese, J. A. Frieman and
A. Olinto, 1993, Phys. Rev. D {\bf 47}, 426.
\item Adams, F. C. and K. Freese, 1995, Phys. Rev. D {\bf 51}, 6722.
\item Albrecht A. and P. J. Steinhardt, 1982, Phys. Rev. Lett. { \bf 48},
1220.
\item Alcock, C. et al., 1993, Nat. {\bf 365}, 621.
\item Allen, B., 1988, Phys. Rev. D {\bf 37}, 2078.
\item Allen, B. and S. Koranda, 1994, Phys. Rev. D {\bf 50}, 3713.
\item Allen, T. J., B. Grinstein, and M. B. Wise, 1987, Phys. Lett. B
{\bf 197}, 66.
\item Arnowitt, R., S. Deser and C. W. Misner, 1962, in Gravitation:
An Introduction to Current Research, ed L. Witten
(John Wiley, New York).
\item Atrio--Barandela, F. and J. Silk, 1994, Phys. Rev. D {\bf 49}, 1126.
\item Aubourg, E. et al., 1993, Nat. {\bf 365}, 623.
\item Bar--Kana, R., 1994, Phys. Rev. D {\bf 50}, 1157.
\item Bardeen, J. M., 1980, Phys. Rev. D {\bf 22}, 1882.
\item Bardeen, J. M., P. J. Steinhardt and M. S. Turner, 1983,
Phys. Rev. D {\bf 28}, 679.
\item Barrow, J. D. and R. A. Matzner, 1977, Mon. Not. R. astr. Soc.
{\bf 181}, 719.
\item Barrow, J. D. and S. Cotsakis, 1988, Phys. Lett. B {\bf 214}, 515.
\item Barrow, J. D. and P. Coles, 1991, Mon. Not. R. astr. Soc.
{\bf 248}, 52.
\item Barrow, J. D., E. J. Copeland and A. R. Liddle, 1992, Phys. Rev.
D {\bf 46}, 645.
\item Barrow, J. D. and A. R. Liddle, 1993, Phys. Rev. D {\bf 47}, R5219
\item Bennett, C. L. et al., 1994, Astrophys. J. {\bf 436}, 423.
\item Bennett, C. L. et al., 1996, Astrophys. J. {\bf 464}, L1.
\item Bergmann, P. G., 1968, Int. J. Theor. Phys. {\bf 1}, 25.
\item Birrell, N. and P. C. W. Davies, 1982, {\em Quantum Fields in
Curved Space} (Cambridge Univ. Press, Cambridge).
\item Bond, J. R., 1992, in Highlights in Astronomy Vol 9, Proc of the
IAU Joint Discussion, ed J. Bergeron.
\item Bond, J. R. and G. Efstathiou, 1987, Mon. Not. R. astr. Soc.
{\bf 226}, 665.
\item Bond, J. R., R. Crittenden, R. L. Davies, G. Efstathiou and
P. J. Steinhardt, 1994, Phys. Rev. Lett. {\bf 72}, 13.
\item Brustein, R. and G. Veneziano, 1994, Phys. Lett. B {\bf 329}, 429.
\item Brustein, R., M. Gasperini, M. Giovannini and G. Veneziano, 1995,
Phys. Lett. B {\bf 361}, 45.
\item Bucher M., A. S. Goldhaber and N. Turok, 1995, Phys. Rev.
D {\bf 52}, 3314.
\item Bucher M., A. S. Goldhaber and N. Turok, 1995, Nucl. Phys. S {\bf 43}
173.
\item Burigana, C., L. Danese and G. de Zotti, 1991, Astr. Astrophys.
{\bf 246} 49.
\item Carr, B. J., 1975, Astrophys. J. {\bf 205}, 1.
\item Carr, B. J. and J. E. Lidsey, 1993, Phys. Rev. D {\bf 48}, 543.
\item Carr, B. J., J. H. Gilbert and J. E. Lidsey, 1994, Phys. Rev.
D {\bf 50}, 4853.
\item Cen, R., N. Y. Gnedin, L. A. Kofman and J. P. Ostriker, 1992,
Astrophys. J. Lett. {\bf 399}, L11.
\item Christensen, N., 1992, Phys. Rev. D {\bf 46}, 5250.
\item Copeland, E. J., E. W. Kolb, A. R. Liddle and J. E. Lidsey, 1993a,
Phys. Rev. Lett. {\bf 71}, 219.
\item Copeland, E. J., E. W. Kolb, A. R. Liddle and J. E. Lidsey, 1993b,
Phys. Rev. D {\bf 48}, 2529 [CKLL1].
\item Copeland, E. J., E. W. Kolb, A. R. Liddle and J. E. Lidsey, 1994a,
Phys. Rev. D {\bf 49}, 1840 [CKLL2].
\item Copeland, E. J., A. R. Liddle, D. H. Lyth, E. D. Stewart and
D. Wands, 1994b, Phys. Rev. D {\bf 49}, 6410.
\item Copi, C. J., D. N. Schramm and M. S. Turner, 1995, Science {\bf 267},
192.
\item Crittenden, R., J. R. Bond, R. L. Davis, G. Efstathiou and P. J.
Steinhardt, 1993a, Phys. Rev. Lett. {\bf 71}, 324.
\item Crittenden, R., R. L. Davis and P. J. Steinhardt, 1993b,
Astrophys. J. Lett. {\bf 417}, L13.
\item Crittenden, R., D. Coulson and N. Turok, 1994, Phys. Rev. Lett.
{\bf 73}, 2390.
\item Daly, R. A., 1991, Astrophys. J. {\bf 371}, 14.
\item Danese, L. and G. de Zotti, 1977, Rivista del Nuovo Cimento
{\bf 7}, 277.
\item Danzmann, K., 1995, Ann. N. Y. Acad. Sci. {\bf 759}, 481.
\item Davis, R. L., H. M. Hodges, G. F. Smoot, P. J. Steinhardt and M. S.
Turner, 1992, Phys. Rev. Lett. {\bf 69}, 1856.
\item Dodelson, S., L. Knox and E. W. Kolb, 1994, Phys. Rev. Lett.
{\bf 72}, 3444.
\item Dodelson, S., E. Gates and A. Stebbins, 1996, Astrophys. J. {\bf
467}, 10.
\item Dolgov, A. and J. Silk, 1993, Phys. Rev. D {\bf 47}, 2619.
\item Durrer, R. and M. Sakellariadou, 1994, Phys. Rev. D {\bf 50}, 6115.
\item Easther, R., 1996, Class. Quant. Grav. {\bf 13}, 1775.
\item Easther, R., K. Maeda and D. Wands, 1996, Phys. Rev. D {\bf 53},
4247.
\item Efstathiou, G. P., 1990, in Physics of the Early Universe, eds
A. T. Davies, A. Heavens and J. Peacock, SUSSP publications
(Edinburgh).
\item Faller, J. E., P. L. Bender, J. L. Hall, D. Hils and M. A. Vincent.,
1985, in Proceedings of the Colloquium Kilometric Optical
Arrays in Space (European Space Agency, Noordwijk).
\item Freedman, W. L. et al., 1994, Nat. {\bf 371}, 757.
\item Frewin, R. A., A. G. Polnarev and P. Coles, 1994, Mon. Not. R.
astr. Soc. {\bf 266}, L21.
\item Garc\'{\i}-Bellido, J. and D. Wands, 1995, Phys. Rev. D {\bf 52},
6739.
\item Garc\'{\i}-Bellido, J. and D. Wands, 1996, Phys. Rev. D {\bf 53},
5437.
\item Garc\'{\i}-Bellido, J., Linde A. D. and D. Wands, 1996,
``Density Perturbations and Black Hole Formation in Hybrid
Inflation'', Phys. Rev. D, to appear, astro-ph/9605094.
\item Gasperini, M. and G. Veneziano, 1993a, Astropart. Phys. {\bf 1},
317.
\item Gasperini, M. and G. Veneziano, 1993b, Mod. Phys. Lett. A {\bf 8},
3701.
\item Gasperini, M. and G. Veneziano, 1994, Phys. Rev. D {\bf 50}, 2519.
\item Gasperini, M., M. Giovannini and G. Veneziano, 1995, Phys. Rev.
D {\bf 52}, R6651.
\item Gott, J. R., 1982, Nature {\bf 295}, 304.
\item Gott, J. R. and T. S. Statler, 1984, Phys. Lett. B {\bf 136}, 157.
\item Grib, A. A., S. Mamaev and V. Mostepanenko, 1980, {\em Quantum
Effects in Strong External Fields} (Atomizdat, Moscow).
\item Grishchuk, L. P., 1974, Zh. Eksp. Teor. Fiz. {\bf 67}, 825 (Sov.
Phys. JETP {\bf 40}, 409).
\item Grishchuk, L. P., 1977, Ann. N. Y. Acad. Sci. {\bf 302}, 439.
\item Grishchuk, L. P., 1989, Sov. Phys. Usp. {\bf 31}, 940.
\item Grishchuk, L. P. and Yu. V. Sidorav, 1988, in Fourth Seminar on
Quantum Gravity, eds M. A. Markov, V. A. Berezin and V. P.
Frolov (World Scientific, Singapore).
\item Grishchuk, L. P. and M. Solokhin, 1991, Phys. Rev. D {\bf 43}, 2566.
\item Guth, A. H., 1981, Phys. Rev. D {\bf 23}, 347.
\item Guth, A. H. and S. -Y. Pi, 1982, Phys. Rev. Lett. {\bf 49}, 1110.
\item Harrison, E. R., 1970, Phys. Rev. D {\bf 1}, 2726.
\item Hawking, S. W. and G. F. R. Ellis, 1973, {\em The Large
Scale Structure of Space--time} (Cambridge University Press,
Cambridge).
\item Hawking, S. W., 1982, Phys. Lett. B {\bf 115}, 295.
\item Hawking, S. W. and I. G. Moss, 1982, Phys. Lett. B {\bf 110}, 35.
\item Higgs, P. W., 1959, Nuovo Cimento {\bf 11}, 816.
\item Hodges, H. M. and G. R. Blumenthal, 1990, Phys. Rev. D {\bf 42},
3329.
\item Hu, W., D. Scott and J. Silk, 1994, Astrophys. J. Lett. {\bf 430}, L5.
\item Hu, W. and N. Sugiyama, 1995, Phys. Rev. D {\bf 51}, 2599.
\item Hu, W., D. Scott, N. Sugiyama and M. White, 1995, Phys. Rev.
D {\bf 52}, 5498.
\item Ivanov, P., P. Nasel'skii and I. Novikov, 1994, Phys. Rev.
D {\bf 50}, 7173.
\item Jungman, G., M. Kamionkowski, A. Kosowsky and D. N. Spergel,
1996, Phys. Rev. D {\bf 54}, 1332.
\item Kalara, S., N. Kaloper and K. A. Olive, 1990, Nucl. Phys. B {\bf
341}, 252.
\item Kaloper, N., R. Madden and K. A. Olive, 1995, Nucl. Phys. B
{\bf 452}, 677.
\item Kaloper, N., R. Madden and K. A. Olive, 1996, Phys. Lett. B
{\bf 371}, 34.
\item Kamionkowski M., Kosowsky A., Stebbins, A., 1996, ``A probe of
primordial gravitational waves and vorticity'', Columbia
preprint, astro-ph/9609132.
\item Khlopov, M. Yu., B. A. Malomed and Ya. B. Zel'dovich, 1985,
Mon. Not. R. astr. Soc. {\bf 215}, 575.
\item Knox, L. and M. S. Turner, 1994, Phys. Rev. Lett. {\bf 73}, 3347.
\item Knox, L., 1995, Phys. Rev. D {\bf 52}, 4307.
\item Kodama, H. and M. Sasaki, 1984, Prog. Theor. Phys. Supp. {\bf 78}, 1.
\item Kolb, E. W. and M. S. Turner, 1990, {\em The Early Universe},
(Addison-Wesley, Redwood City, California).
\item Kolb, E. W. and S. L. Vadas, 1994, Phys. Rev. D {\bf 50}, 2479.
\item Koranda, S. and B. Allen, 1994, Phys. Rev. D {\bf 50}, 3713.
\item Kosowsky, A., 1996, Annals. Phys. {\bf 246}, 49.
\item Kosowsky, A. and M. S. Turner, 1995, Phys. Rev. D {\bf 52}, R1739.
\item Krauss, L. M. and M. White, 1992, Phys. Rev. Lett. {\bf 69}, 869.
\item Levin, J. J., 1995a, Phys. Rev. D {\bf 51}, 462.
\item Levin, J. J. 1995b, Phys. Rev. D {\bf 51}, 1536.
\item Liddle, A. R., 1994a, Phys. Rev. D {\bf 49}, 739.
\item Liddle, A. R., 1994b, Phys. Rev. D {\bf 49}, 3805; Phys. Rev.
D {\bf 51}, 4603 (E).
\item Liddle, A. R. and D. H. Lyth, 1992, Phys. Lett. B {\bf 291}, 391.
\item Liddle, A. R. and D. H. Lyth, 1993a, Phys. Rept. {\bf 231}, 1.
\item Liddle, A. R. and D. H. Lyth, 1993b, Mon. Not. R. astr. Soc.
{\bf 265}, 379.
\item Liddle, A. R., D. H. Lyth and W. Sutherland, 1992, Phys. Lett. B
{\bf 279}, 244.
\item Liddle, A. R. and M. S. Turner, 1994, Phys. Rev. D {\bf 50}, 758.
\item Liddle, A. R., P. Parsons and J. D. Barrow, 1994, Phys. Rev.
D {\bf 50}, 7222.
\item Liddle, A. R., D. H. Lyth, R. K. Schaefer, Q. Shafi and P. T. V.
Viana, 1996, Mon. Not. R. astr. Soc. {\bf 281}, 531.
\item Lidsey, J. E., 1991a, Class. Quant. Grav. {\bf 8}, 923.
\item Lidsey, J. E., 1991b, Phys. Lett. B {\bf 273}, 42.
\item Lidsey, J. E., 1992, Class. Quant. Grav. {\bf 9}, 149.
\item Lidsey, J. E., 1993, Gen. Rel. Grav. {\bf 25}, 399.
\item Lidsey, J. E. and P. Coles, 1992, Mon. Not. R. astr. Soc.
{\bf 258}, 57P.
\item Lifshitz, E. M., 1946, Zh. Eksp. Teor. Phys. {\bf 16}, 587.
\item Linde, A. D., 1982a, Phys. Lett. B {\bf 108}, 389.
\item Linde, A. D., 1982b, Phys. Lett. B {\bf 116}, 335.
\item Linde, A. D., 1983, Phys. Lett. B {\bf 129}, 177.
\item Linde, A. D., 1990a, Phys. Lett. B {\bf 249}, 18.
\item Linde, A. D., 1990b, {\em Particle Physics and Inflationary
Cosmology} (Harwood Academic, Chur, Switzerland).
\item Linde, A. D., 1991, Phys. Lett. B {\bf 259}, 38.
\item Linde, A. D., 1994, Phys. Rev. D {\bf 49}, 748.
\item Linde, A. D., 1995, Phys. Lett. B {\bf 351}, 99.
\item Linde, A. D. and A. Mezhlumian, 1995, Phys. Rev. D {\bf 52}, 6789.
\item Lindley, D., 1980, Mon. Not. R. astr. Soc. {\bf 193}, 593.
\item Lucchin, F. and S. Matarrese, 1985a, Phys. Rev. D {\bf 32}, 1316.
\item Lucchin, F. and S. Matarrese, 1985b, Phys. Lett. B {\bf 164}, 282.
\item Lucchin, F., S. Matarrese and S. Mollerach, 1992, Astrophys. J.
Lett. {\bf 401}, 49.
\item Lyth, D. H., 1985, Phys. Rev. D {\bf 31}, 1792.
\item Lyth, D. H. and E. D. Stewart, 1992, Phys. Lett. B {\bf 274}, 168.
\item Lyth, D. H., 1996, ``What would we learn by detecting a gravitational
wave signal in the cosmic microwave background anisotropy?'',
Lancaster preprint, hep-ph/9606387.
\item Ma, C.-P. and E. Bertschinger, 1995, Astrophys. J. {\bf 455}, 7.
\item MacGibbon, J. H., 1987, Nature 320, 308.
\item MacGibbon, J. H. and B. J. Carr, 1991, Astrophys. J. {\bf 371}, 447.
\item Maeda, K., 1989, Phys. Rev. D {\bf 39}, 3159.
\item Makino, N. and M. Sasaki, 1991, Prog. Theor. Phys. {\bf 86}, 103.
\item Mangano, G., G. Miele and C. Stornaiolo, 1995, Mod. Phys. Lett. A
{\bf 10}, 1977.
\item Mather, J. C. et al., 1994, Astrophys. J. {\bf 420}, 439.
\item Mielke, E. W. and F. E. Schunck, 1995, Phys. Rev. D {\bf 52}, 672.
\item Misner, C. W., K. S. Thorne and J. A. Wheeler, 1973,
{\em Gravitation} (Freeman, San Francisco).
\item Mollerach, S., S. Matarrese and F. Lucchin, 1994, Phys. Rev.
D {\bf 50}, 4835.
\item Mukhanov, V. F., 1985, Pis'ma. Zh. Eksp. Teor. Fiz. {\bf 41},
402 [JETP Lett. {\bf 41}, 493].
\item Mukhanov, V. F., 1988, Zh. Eksp. Teor. Fiz. {\bf 94}, 1 [Sov.
Phys. JETP {\bf 41}, 493].
\item Mukhanov, V. F., 1989, Phys. Lett. B {\bf 218}, 17.
\item Mukhanov, V. F., H. A. Feldman and R. H. Brandenberger, 1992,
Phys. Rept. {\bf 215}, 203.
\item Muslimov, A. G., 1990, Class. Quant. Grav. {\bf 7}, 231.
\item Nakamura, T. T. and E. D. Stewart, 1996, Phys. Lett. B {\bf 381},
413.
\item Nasel'skii, P. D. and A. G. Polnarev, 1985, Astron. Zh. {\bf 62},
833 (Sov. Astron. {\bf 29}, 487).
\item Olive, K. A., 1990, Phys. Rept. {\bf 190}, 307.
\item Olive, K. A., D. N. Schramm, G. Steigman and T. Walker, 1990,
Phys. Lett. B {\bf 236}, 454.
\item Peebles, P. J. E. and R. J. Dicke, 1979, in General Relativity,
eds S. W. Hawking and W. Israel (Cambridge University Press,
Cambridge).
\item Pogosyan, D. Yu. and A. A. Starobinsky, 1995, Astrophys. J.
{\bf 447}, 465.
\item Polarski D. and A. A. Starobinsky, 1995, Phys. Lett. B {\bf 356},
196.
\item Pollock, M. D. and D. Sahdev, 1989, Phys. Lett. B {\bf 222}, 12.
\item Randall, L., M. Solja\u{c}i\'{c} and A. H. Guth., 1996,
Nucl. Phys. B {\bf 472}, 377.
\item Sachs, R. K. and A. M. Wolfe, 1967, Astrophys. J. {\bf 147}, 73.
\item Sahni, V., 1990, Phys. Rev. D {\bf 42}, 453.
\item Salopek, D. S., 1992, Phys. Rev. Lett. {\bf 69}, 3602.
\item Salopek, D. S, J. R. Bond and J. M. Bardeen, 1989, Phys. Rev.
D {\bf 40}, 1753.
\item Salopek, D. S. and J. R. Bond, 1990, Phys. Rev. D {\bf 42}, 3936.
\item Salopek, D. S. and J. R. Bond, 1991, Phys. Rev. D {\bf 43}, 1005.
\item Sasaki, M., 1986, Prog. Theor. Phys. {\bf 76}, 1036.
\item Sasaki, M., T. Tanaka, K. Yamamoto and J. Yokoyama, 1993,
Phys. Lett. B {\bf 317}, 510.
\item Sasaki, M. and E. D. Stewart, 1996, Prog. Theor. Phys. {\bf 95}, 71.
\item Sato, K., 1981, Mon. Not. R. astr. Soc. {\bf 195}, 467.
\item Schaefer, R. K. and Q. Shafi, 1994, Phys. Rev. D {\bf 49}, 4990.
\item Seljak, U. and E. Bertschinger, 1994, in ``Present and Future of
the Cosmic Microwave Background'', eds. J.L. Sanz, E.
Martinez-Gonzalez and L. Cajon. Springer Verlag (Berlin).
\item Seljak, U. and Zaldarriaga, M., 1996a, Astrophys. J. {\bf 469}, 437.
\item Seljak, U. and Zaldarriaga, M., 1996b, ``Signature of gravity
waves in polarization of the microwave background'', CfA preprint,
astro-ph/9609169.
\item Silk, J., 1967, Astrophys. J. {\bf 151}, 459.
\item Smoot, G. F. et al., 1992, Astrophys. J. Lett. {\bf 396}, L1.
\item Souradeep, T. and V. Sahni, 1992, Mod. Phys. Lett. A {\bf 7}, 3541.
\item Starobinsky, A. A., 1979, Pis'ma Zh. Eksp. Teor. Fiz. {\bf 30},
719 (JETP Letters {\bf 30}, 682).
\item Starobinsky, A. A., 1980, Phys. Lett. B {\bf 91}, 99.
\item Starobinsky, A. A., 1982 Phys. Lett. B {\bf 117}, 175.
\item Starobinsky, A. A., 1985, Pis'ma Astron. Zh. {\bf 11}, 323 (Sov.
Astron. Lett. {\bf 11}, 133).
\item Starobinsky, A. A. and J. Yokoyama, 1995, ``Density
Perturbations in Brans--Dicke theory'', Kyoto preprint,
gr-qc/9502002.
\item Stebbins, R. T., P. L. Bender, J. E. Faller, J. L. Hall, D. Hils,
and M. A. Vincent., 1989, in Fifth Marcel Grossman Meeting
(World Scientific, Singapore).
\item Steinhardt, P. J. and M. S. Turner, 1984, Phys. Rev. D {\bf 29}, 2162.
\item Steinhardt, P. J., 1994, in ``Anisotropies two years after
COBE'', ed L. M. Krauss (World Scientific, Singapore).
\item Stewart, E. D. and D. H. Lyth, 1993, Phys. Lett. B {\bf 302}, 171.
\item Stompor, R., 1994, Astron. Astrophys. {\bf 287}, 693
\item Sunyaev, R. A. and Ya. B. Zel'dovich, 1970, Astrophys. Sp. Sci.
{\bf 7}, 20.
\item Sunyaev, R. A. and Ya. B. Zel'dovich, 1980, Ann. Rev. Astr. Astrophys.
{\bf 18}, 537.
\item Tegmark, M. and G. Efstathiou, 1996, Mon. Not. R. astr. Soc.
{\bf 281}, 1297.
\item Thorne, K. S., 1987, in 300 years of Gravitation, eds
S. W. Hawking and W. Israel (Cambridge Univ. Press, Cambridge).
\item Thorne, K. S., 1995, ``Gravitational Waves'', CalTech preprint,
gr-qc/9506086.
\item Turner, M. S., 1993a, Phys. Rev. D {\bf 48}, 3502.
\item Turner, M. S., 1993b, Phys. Rev. D {\bf 48}, 5539.
\item Turner, M. S., M. White and J. E. Lidsey, 1993, Phys. Rev.
D {\bf 48}, 4613.
\item Turner, M. S. and M. White, 1996, Phys. Rev. D {\bf 53}, 6822.
\item Vilenkin, A. and Shellard, E. P. S., 1994, {\em Cosmic Strings and
other topological defects} (Cambridge University Press).
\item Wagoner, R. V., 1970, Phys. Rev. D {\bf 1}, 3204.
\item Wands, D., 1994, Class. Quant. Grav. {\bf 11}, 269.
\item Wang, Y., 1996, Phys. Rev. D {\bf 53}, 639.
\item White, M., 1992, Phys. Rev. D {\bf 42}, 4198.
\item White, M., D. Scott and J. Silk, 1994, Ann. Rev. Astron. Astrophys.
{\bf 32}, 319.
\item White, M., D. Scott, J. Silk and M. Davis, 1995, Mon. Not. R.
astr. Soc. {\bf 276}, L69.
\item Whitt, B., 1984, Phys. Lett. B {\bf 145}, 176.
\item Wright, E. et al., 1992, Astrophys. J. Lett. {\bf 396}, L13.
\item Yokoyama, J., 1995, ``Formation of MACHO-primordial black holes
in inflationary cosmology'', YITP preprint.
\item Zaldarriaga, M. and Seljak, U., 1996, ``An all-sky analysis of
polarization in the microwave background'', CfA preprint,
astro-ph/9609170
\item Zel'dovich, Ya. B., 1972, Mon. Not. R. astr. Soc. {\bf 160}, 1P.
\item Zel'dovich, Ya. B. and A. A. Starobinsky, 1976, Pis'ma Zh. Eksp. Teor.
Fiz. {\bf 24}, 616 (1976) (JETP Lett. {\bf 24}, 571).
\end{description}
\newpage
\section*{Tables}
\begin{table}[h]
\begin{center}
\begin{tabular}{c|c|c}
\hline \hline
& Gravitational Waves & Gravitational Waves \\
& Important & Negligible \\
\hline \hline
& & \\
$n<1$ & $\epsilon$ large, $\eta < 2\epsilon$ & $\epsilon$ small, $\eta <
-2\epsilon$ \\
& Power-Law Inflation & Natural Inflation \\
\hline
$n \simeq 1$ & $\epsilon$ large, $\eta \simeq 2\epsilon$ & $\epsilon$,
$|\eta|$ small \\
& Intermediate Inflation & Hybrid Inflation \\
\hline
$n>1$ & $\epsilon$ large, $\eta > 2\epsilon$ & $\epsilon$ small, $\eta >
2\epsilon$ \\
& & Hybrid Inflation\\
\end{tabular}
\end{center}
\footnotesize{\hspace*{.3in} Table 1: This table illustrates the different
possible inflationary behaviors, and quotes a specific inflation
model which gives each (except the bottom left case, which while
possible in principle has not had any specific inflationary model
devised). The description `large' implies significantly larger than
zero (but still less than unity).}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{c|c|c}
\hline \hline
& lowest-order & next-order (exact) \\
\hline \hline
& & \\
$V(\phi_0)$ & $H(\phi_0)$
& $H(\phi_0)$, $\epsilon(\phi_0)$ \\
$V'(\phi_0)$ & $H(\phi_0)$, $\epsilon(\phi_0)$
& $H(\phi_0)$, $\epsilon(\phi_0)$, $\eta(\phi_0)$ \\
$V''(\phi_0)$ & $H(\phi_0)$, $\epsilon(\phi_0)$, $\eta(\phi_0)$
& $H(\phi_0)$, $\epsilon(\phi_0)$, $\eta(\phi_0)$, $\xi(\phi_0)$ \\
$V'''(\phi_0)$ & $H$, $\epsilon(\phi_0)$, $\eta(\phi_0)$, $\xi(\phi_0)$
& -------- \\
\end{tabular}
\end{center}
\footnotesize {\hspace*{.3in} Table 2: A summary of the inflationary
parameters [$H$ and the slow-roll parameters $\epsilon$, $\eta$, and
$\xi$ defined in Eqs.\ (\ref{epsilon})--(\ref{xi})] needed to
reconstruct a given derivative of the potential to a certain order.
See Eqs.\ (\ref{V})--(\ref{V''}). Note that the next-order result is {\em
exact}.}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{c|c|c}
\hline \hline
observable & lowest-order & next-order \\
\hline \hline
& & \\
$A_T^2(k_0)$ & $H(\phi_0)$
& $H(\phi_0)$, $\epsilon(\phi_0)$ \\
$A_S^2(k_0)$ & $H(\phi_0)$, $\epsilon(\phi_0)$ &
$H(\phi_0)$, $\epsilon(\phi_0)$, $\eta(\phi_0)$ \\
$n(k_0)$ & $\epsilon(\phi_0)$, $\eta(\phi_0)$ &
$\epsilon(\phi_0)$, $\eta(\phi_0)$, $\xi(\phi_0)$\\
$dn/d\ln k|_{k_0}$ & $\epsilon(\phi_0)$, $\eta(\phi_0)$, $\xi(\phi_0)$
& -------- \\
\end{tabular}
\end{center}
\footnotesize {\hspace*{.3in} Table 3: The observables, $A_T^2$, $A_S^2$,
$n$, and $dn/d\ln k$ at the point $k_0$ may be expressed in terms of $H$ and
the slow-roll parameters at the point $\phi_0$. Table 3 lists the inflation
parameters
required to predict the observable to the indicated order. (See Section
5.1.)}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{c|c|c}
\hline \hline
parameter & lowest-order & next-order \\
\hline \hline
& & \\
$H$
& $A_T^2$
& $A_T^2$, $A_S^2$ \\
$\epsilon$
& $A_T^2$, $A_S^2$
& $A_T^2$, $A_S^2$, $n$ \\
$\eta$
& $A_T^2$, $A_S^2$, $n$
& $A_T^2$, $A_S^2$, $n$, $dn/d\ln k$ \\
$\xi$
& $A_T^2$, $A_S^2$, $n$, $dn/d\ln k$
& -------- \
\end{tabular}
\end{center}
\footnotesize {\hspace*{.3in} Table 4: The inflation parameters
may be expressed in terms of the observables, $A_T^2$, $A_S^2$,
$n$, and $dn/d\ln k$ (see Section 5.1). Through judicious use of the
consistency relations one may employ different combinations of observables
than listed here, e.g., use of $n_T$ rather than $A_T^2/A_S^2$.}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{c|c|c|c}
\hline \hline
& lowest-order & next-order & next-to-next-order \\
\hline \hline
& & &\\
$V$
& $A_T^2$
& $A_T^2$, $A_S^2$
& $A_T^2$, $A_S^2$, $n$ \\
$V'$
& $A_T^2$, $A_S^2$
& $A_T^2$, $A_S^2$, $n$
& $A_T^2$, $A_S^2$, $n$, $dn/d\ln k$ \\
$V''$
& $A_T^2$, $A_S^2$, $n$
& $A_T^2$, $A_S^2$, $n$, $dn/d\ln k$
& -------- \\
$V'''$
& $A_T^2$, $A_S^2$, $n$, $dn/d\ln k$
& --------
& -------- \\
\end{tabular}
\end{center}
\footnotesize {\hspace*{.3in} Table 5: A summary of the observables
needed to reconstruct a given derivative of the potential to a certain
order. The potential and its derivatives are given at a point $\phi_0$, and
$A_T^2$, $A_S^2$, $n$, and $dn/dk$ are to be evaluated at the point $k_0$.}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{c|c|c|c}
\hline \hline
Model 1 & Input & Output (power-law fit) & Output
(including $dn/d\ln k|_{k_0}$) \\
\hline \hline
& & &\\
$A_S^2$ & $2.5 \times 10^{-10}$ & $(2.45 \pm 0.09) \times 10^{-10}$ &
$(2.45 \pm 0.10) \times 10^{-10}$ \\
$A_T^2$ & $0.12 \times 10^{-10}$ & $(0.132 \pm 0.015) \times 10^{-10}$ &
$(0.132 \pm 0.015) \times 10^{-10}$ \\
$n-1$ & $-0.1$ & $-0.11 \pm 0.02$ & $-0.115 \pm 0.035$\\
$n_T$ & $-0.1$ & $-0.25 \pm 0.10$ & $-0.25 \pm 0.10$\\
$dn/d\ln k|_{k_0}$ & $0$ & --- & $0.003 \pm 0.018$
\end{tabular}
\vspace*{12pt}
\begin{tabular}{c|c|c|c}
\hline \hline
Model 2 & Input & Output (power-law fit) & Output
(including $dn/d\ln k|_{k_0}$) \\
\hline \hline
& & &\\
$A_S^2$ & $1.34 \times 10^{-10}$ & $(1.27 \pm 0.04) \times 10^{-10}$ &
$(1.28 \pm 0.04) \times 10^{-10}$ \\
$A_T^2$ & $0.094 \times 10^{-10}$ & $(0.09 \pm 0.01) \times 10^{-10}$ &
$(0.09 \pm 0.01) \times 10^{-10}$ \\
$n-1$ & $0.00$ & $0.04 \pm 0.02$ & $0.06\pm 0.03$\\
$n_T$ & $-0.2$ & $-0.12 \pm 0.11$ & $-0.12 \pm 0.11$\\
$dn/d\ln k|_{k_0}$ & $0$ & --- & $-0.01 \pm 0.02$
\end{tabular}
\end{center}
\footnotesize {\hspace*{.3in} Table 6: Input and output values from the
two simulated data sets. The amplitudes are given at the central $k$
value (in log units) for the scalars, notionally corresponding to the
20-th multipole.}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{c|c|c|c}
\hline \hline
Model 1 & Underlying & Lowest-order & Next-order \\
& potential & reconstruction & reconstruction \\
\hline \hline
& & &\\
$10^{12} V(\phi_0)/m_{{\rm Pl}}^4$ & 28.2 & $31 \pm 4$ & $31 \pm 4$ \\
$10^{12} V'(\phi_0)/m_{{\rm Pl}}^3$ & -43.6 & $-51 \pm 9$ &
$-52 \pm 9$ \\
$10^{12} V''(\phi_0)/m_{{\rm Pl}}^2$ & 67.5 & $83 \pm 25$ & ---
\end{tabular}
\vspace*{12pt}
\begin{tabular}{c|c|c|c}
\hline \hline
Model 2 & Underlying & Lowest-order & Next-order \\
& potential & reconstruction & reconstruction \\
\hline \hline
& & &\\
$10^{12} V(\phi_0)/m_{{\rm Pl}}^4$ & 22.4 & $21 \pm 2$ & $21\pm 2$ \\
$10^{12} V'(\phi_0)/m_{{\rm Pl}}^3$ & -38.9 & $-40 \pm 7$ & $-37 \pm 6$ \\
$10^{12} V''(\phi_0)/m_{{\rm Pl}}^2$ & 94.5 & $123 \pm 27$ & ---
\end{tabular}
\end{center}
\footnotesize {\hspace*{.3in} Table 7: Input potential compared with
reconstructions for the two models.}
\end{table}
\newpage
\section*{Figure Captions}
\vspace*{24pt}
\noindent
{\em Figure 1}\\
A schematic illustration of the reconstruction strategy. The spectra $A_S$
of the density perturbations and $A_T$ of the gravitational waves are
measured over some range of scales which corresponds to some interval of the
underlying potential $V(\phi)$.
\vspace*{24pt}
\noindent
{\em Figure 2}\\
The simulated data of Model 1, with error bars. The circles are $A_S^2$ and
squares are $A_T^2$. The horizontal axis is in $h \, {\rm Mpc}^{-1}$. The
lines show the best power-law fits to the simulated data, as given in Table
2. Showing the data in the form of the spectra is schematic; an analysis of
true observations would directly fit the amplitude and spectral index to
measured quantities.
\vspace*{24pt}
\noindent
{\em Figure 3}\\
The reconstructed potentials compared to the underlying one, from the data
in Model 1 in Table 4. The dashed line shows the true underlying exponential
potential. The two solid lines, which nearly overlap, are Taylor series
reconstructions, one using just lowest-order information and the other using
the available next-order information. The length of these lines corresponds
to the range of $k$ for which the simulated data is available. The
observational errors (not shown) dominate the theoretical errors, and of
course when taken into account the reconstructions are consistent with the
true potential.
\vspace*{24pt}
\noindent
{\em Figure 4}\\
An illustration of the area law. Reconstruction finds $\epsilon$ and
perhaps its derivative, between 60 and 50 $e$-foldings from the end of
inflation, illustrated by the solid part of the curve which ends at a scalar
field value indicated by $\phi_{50}$. After large-scale structure scales
leave the horizon, $\epsilon$ (now shown as a dotted curve) must behave so
that it reaches unity just as the shaded area under the curve of
$\epsilon^{-1/2}$ against $\phi/m_{{\rm Pl}}$ reaches $50/\sqrt{4\pi}$.
\end{document}
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.